23.11.2012 Views

Contact Geometry - Mathematisches Institut der Universität zu Köln

Contact Geometry - Mathematisches Institut der Universität zu Köln

Contact Geometry - Mathematisches Institut der Universität zu Köln

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Contents<br />

<strong>Contact</strong> <strong>Geometry</strong><br />

Hansjörg Geiges<br />

<strong>Mathematisches</strong> <strong>Institut</strong>, <strong>Universität</strong> <strong>zu</strong> <strong>Köln</strong>,<br />

Weyertal 86–90, 50931 <strong>Köln</strong>, Germany<br />

E-mail: geiges@math.uni-koeln.de<br />

April 2004<br />

1 Introduction 3<br />

2 <strong>Contact</strong> manifolds 4<br />

2.1 <strong>Contact</strong> manifolds and their submanifolds . . . . . . . . . . . . . . 6<br />

2.2 Gray stability and the Moser trick . . . . . . . . . . . . . . . . . . 13<br />

2.3 <strong>Contact</strong> Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.4 Darboux’s theorem and neighbourhood theorems . . . . . . . . . . 17<br />

2.4.1 Darboux’s theorem . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.4.2 Isotropic submanifolds . . . . . . . . . . . . . . . . . . . . . 19<br />

2.4.3 <strong>Contact</strong> submanifolds . . . . . . . . . . . . . . . . . . . . . 24<br />

2.4.4 Hypersurfaces . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.4.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

2.5 Isotopy extension theorems . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.5.1 Isotropic submanifolds . . . . . . . . . . . . . . . . . . . . . 32<br />

2.5.2 <strong>Contact</strong> submanifolds . . . . . . . . . . . . . . . . . . . . . 34<br />

2.5.3 Surfaces in 3–manifolds . . . . . . . . . . . . . . . . . . . . 36<br />

2.6 Approximation theorems . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.6.1 Legendrian knots . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.6.2 Transverse knots . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

1


3 <strong>Contact</strong> structures on 3–manifolds 43<br />

3.1 An invariant of transverse knots . . . . . . . . . . . . . . . . . . . . 45<br />

3.2 Martinet’s construction . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.3 2–plane fields on 3–manifolds . . . . . . . . . . . . . . . . . . . . . 50<br />

3.3.1 Hopf’s Umkehrhomomorphismus . . . . . . . . . . . . . . . 53<br />

3.3.2 Representing homology classes by submanifolds . . . . . . . 54<br />

3.3.3 Framed cobordisms . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.3.4 Definition of the obstruction classes . . . . . . . . . . . . . 57<br />

3.4 Let’s Twist Again . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.5 Other existence proofs . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.5.1 Open books . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.5.2 Branched covers . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

3.5.3 . . . and more . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.6 Tight and overtwisted . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.7 Classification results . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

4 A guide to the literature 75<br />

4.1 Dimension 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.2 Higher dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.3 Symplectic fillings . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.4 Dynamics of the Reeb vector field . . . . . . . . . . . . . . . . . . . 77<br />

2


1 Introduction<br />

Over the past two decades, contact geometry has un<strong>der</strong>gone a veritable meta-<br />

morphosis: once the ugly duckling known as ‘the odd-dimensional analogue of<br />

symplectic geometry’, it has now evolved into a proud field of study in its own<br />

right. As is typical for a period of rapid development in an area of mathematics,<br />

there are a fair number of folklore results that every mathematician working in<br />

the area knows, but no references that make these results accessible to the novice.<br />

I therefore take the present article as an opportunity to take stock of some of that<br />

folklore.<br />

There are many excellent surveys covering specific aspects of contact geometry<br />

(e.g. classification questions in dimension 3, dynamics of the Reeb vector field,<br />

various notions of symplectic fillability, transverse and Legendrian knots and<br />

links). All these topics deserve to be included in a comprehensive survey, but<br />

an attempt to do so here would have left this article in the ‘to appear’ limbo for<br />

much too long.<br />

Thus, instead of adding yet another survey, my plan here is to cover in detail<br />

some of the more fundamental differential topological aspects of contact geometry.<br />

In doing so, I have not tried to hide my own idiosyncrasies and preoccupations.<br />

Owing to a relatively leisurely pace and constraints of the present format, I<br />

have not been able to cover quite as much material as I should have wished.<br />

Nonetheless, I hope that the rea<strong>der</strong> of the present handbook chapter will be<br />

better prepared to study some of the surveys I alluded to – a guide to these<br />

surveys will be provided – and from there to move on to the original literature.<br />

A book chapter with comparable aims is Chapter 8 in [1]. It seemed opportune<br />

to be brief on topics that are covered extensively there, even if it is done at the<br />

cost of leaving out some essential issues. I hope to return to the material of the<br />

present chapter in a yet to be written more comprehensive monograph.<br />

Acknowledgements. I am grateful to Fan Ding, Jesús Gonzalo and Fe<strong>der</strong>ica<br />

Pasquotto for their attentive reading of the original manuscript. I also thank<br />

John Etnyre and Stephan Schönenberger for allowing me to use a couple of their<br />

figures (viz., Figures 2 and 1 of the present text, respectively).<br />

3


2 <strong>Contact</strong> manifolds<br />

Let M be a differential manifold and ξ ⊂ TM a field of hyperplanes on M. Locally<br />

such a hyperplane field can always be written as the kernel of a non-vanishing<br />

1–form α. One way to see this is to choose an auxiliary Riemannian metric g on<br />

M and then to define α = g(X, .), where X is a local non-zero section of the line<br />

bundle ξ ⊥ (the orthogonal complement of ξ in TM). We see that the existence<br />

of a globally defined 1–form α with ξ = kerα is equivalent to the orientability<br />

(hence triviality) of ξ ⊥ , i.e. the coorientability of ξ. Except for an example below,<br />

I shall always assume this condition.<br />

If α satisfies the Frobenius integrability condition<br />

α ∧ dα = 0,<br />

then ξ is an integrable hyperplane field (and vice versa), and its integral sub-<br />

manifolds form a codimension 1 foliation of M. Equivalently, this integrability<br />

condition can be written as<br />

X, Y ∈ ξ =⇒ [X, Y ] ∈ ξ.<br />

An integrable hyperplane field is locally of the form dz = 0, where z is a coordi-<br />

nate function on M. Much is known, too, about the global topology of foliations,<br />

cf. [100].<br />

<strong>Contact</strong> structures are in a certain sense the exact opposite of integrable<br />

hyperplane fields.<br />

Definition 2.1. Let M be a manifold of odd dimension 2n + 1. A contact<br />

structure is a maximally non-integrable hyperplane field ξ = kerα ⊂ TM, that<br />

is, the defining 1–form α is required to satisfy<br />

α ∧ (dα) n �= 0<br />

(meaning that it vanishes nowhere). Such a 1–form α is called a contact form.<br />

The pair (M, ξ) is called a contact manifold.<br />

Remark 2.2. Observe that in this case α ∧ (dα) n is a volume form on M; in<br />

particular, M needs to be orientable. The condition α∧(dα) n �= 0 is independent<br />

of the specific choice of α and thus is indeed a property of ξ = kerα: Any other 1–<br />

form defining the same hyperplane field must be of the form λα for some smooth<br />

4


function λ: M → R \ {0}, and we have<br />

(λα) ∧ (d(λα)) n = λα ∧ (λ dα + dλ ∧ α) n = λ n+1 α ∧ (dα) n �= 0.<br />

We see that if n is odd, the sign of this volume form depends only on ξ, not<br />

the choice of α. This makes it possible, given an orientation of M, to speak of<br />

positive and negative contact structures.<br />

Remark 2.3. An equivalent formulation of the contact condition is that we<br />

have (dα) n |ξ �= 0. In particular, for every point p ∈ M, the 2n–dimensional<br />

subspace ξp ⊂ TpM is a vector space on which dα defines a skew-symmetric form<br />

of maximal rank, that is, (ξp, dα|ξp ) is a symplectic vector space. A consequence<br />

of this fact is that there exists a complex bundle structure J : ξ → ξ compatible<br />

with dα (see [92, Prop. 2.63]), i.e. a bundle endomorphism satisfying<br />

• J 2 = −idξ,<br />

• dα(JX, JY ) = dα(X, Y ) for all X, Y ∈ ξ,<br />

• dα(X, JX) > 0 for 0 �= X ∈ ξ.<br />

Remark 2.4. The name ‘contact structure’ has its origins in the fact that one of<br />

the first historical sources of contact manifolds are the so-called spaces of contact<br />

elements (which in fact have to do with ‘contact’ in the differential geometric<br />

sense), see [7] and [45].<br />

In the 3–dimensional case the contact condition can also be formulated as<br />

X, Y ∈ ξ linearly independent =⇒ [X, Y ] �∈ ξ;<br />

this follows immediately from the equation<br />

dα(X, Y ) = X(α(Y )) − Y (α(X)) − α([X, Y ])<br />

and the fact that the contact condition (in dim. 3) may be written as dα|ξ �= 0.<br />

In the present article I shall take it for granted that contact structures are<br />

worthwhile objects of study. As I hope to illustrate, this is fully justified by<br />

the beautiful mathematics to which they have given rise. For an apology of<br />

contact structures in terms of their origin (with hindsight) in physics and the<br />

multifarious connections with other areas of mathematics I refer the rea<strong>der</strong> to the<br />

5


historical surveys [87] and [45]. <strong>Contact</strong> structures may also be justified on the<br />

grounds that they are generic objects: A generic 1–form α on an odd-dimensional<br />

manifold satisfies the contact condition outside a smooth hypersurface, see [89].<br />

Similarly, a generic 1–form α on a 2n–dimensional manifold satisfies the condition<br />

α ∧ (dα) n−1 �= 0 outside a submanifold of codimension 3; such ‘even-contact<br />

manifolds’ have been studied in [51], for instance, but on the whole their theory<br />

is not as rich or well-motivated as that of contact structures.<br />

Definition 2.5. Associated with a contact form α one has the so-called Reeb<br />

vector field Rα, defined by the equations<br />

(i) dα(Rα, .) ≡ 0,<br />

(ii) α(Rα) ≡ 1.<br />

As a skew-symmetric form of maximal rank 2n, the form dα|TpM has a 1–<br />

dimensional kernel for each p ∈ M 2n+1 . Hence equation (i) defines a unique<br />

line field 〈Rα〉 on M. The contact condition α ∧ (dα) n �= 0 implies that α is<br />

non-trivial on that line field, so a global vector field is defined by the additional<br />

normalisation condition (ii).<br />

2.1 <strong>Contact</strong> manifolds and their submanifolds<br />

We begin with some examples of contact manifolds; the simple verification that<br />

the listed 1–forms are contact forms is left to the rea<strong>der</strong>.<br />

Example 2.6. On R 2n+1 with cartesian coordinates (x1, y1, . . .,xn, yn, z), the<br />

1–form<br />

is a contact form.<br />

α1 = dz +<br />

n�<br />

j=1<br />

xj dyj<br />

Example 2.7. On R 2n+1 with polar coordinates (rj, ϕj) for the (xj, yj)–plane,<br />

j = 1, . . .,n, the 1–form<br />

is a contact form.<br />

α2 = dz +<br />

n�<br />

j=1<br />

r 2 j dϕj = dz +<br />

6<br />

n�<br />

(xj dyj − yj dxj)<br />

j=1


x<br />

z<br />

Figure 1: The contact structure ker(dz + x dy).<br />

Definition 2.8. Two contact manifolds (M1, ξ1) and (M2, ξ2) are called contac-<br />

tomorphic if there is a diffeomorphism f : M1 → M2 with Tf(ξ1) = ξ2, where<br />

Tf : TM1 → TM2 denotes the differential of f. If ξi = kerαi, i = 1, 2, this<br />

is equivalent to the existence of a nowhere zero function λ: M1 → R such that<br />

f ∗ α2 = λα1.<br />

Example 2.9. The contact manifolds (R 2n+1 , ξi = kerαi), i = 1, 2, from the<br />

preceding examples are contactomorphic. An explicit contactomorphism f with<br />

f ∗ α2 = α1 is given by<br />

f(x, y, z) = � (x + y)/2, (y − x)/2, z + xy/2 � ,<br />

where x and y stand for (x1, . . .,xn) and (y1, . . .,yn), respectively, and xy stands<br />

for �<br />

j xjyj. Similarly, both these contact structures are contactomorphic to<br />

ker(dz − �<br />

j yj dxj). Any of these contact structures is called the standard<br />

contact structure on R 2n+1 .<br />

Example 2.10. The standard contact structure on the unit sphere S 2n+1<br />

in R 2n+2 (with cartesian coordinates (x1, y1, . . .,xn+1, yn+1)) is defined by the<br />

contact form<br />

n+1 �<br />

α0 = (xj dyj − yj dxj).<br />

j=1<br />

With r denoting the radial coordinate on R2n+2 (that is, r2 = �<br />

j (x2j + y2 j )) one<br />

checks easily that α0 ∧ (dα0) n ∧ r dr �= 0 for r �= 0. Since S2n+1 is a level surface<br />

of r (or r 2 ), this verifies the contact condition.<br />

7<br />

y


Alternatively, one may regard S 2n+1 as the unit sphere in C n+1 with complex<br />

structure J (corresponding to complex coordinates zj = xj+iyj, j = 1, . . .,n+1).<br />

Then ξ0 = kerα0 defines at each point p ∈ S 2n+1 the complex (i.e. J–invariant)<br />

subspace of TpS 2n+1 , that is,<br />

ξ0 = TS 2n+1 ∩ J(TS 2n+1 ).<br />

This follows from the observation that α = −r dr◦J. The hermitian form dα(., J.)<br />

on ξ0 is called the Levi form of the hypersurface S 2n+1 ⊂ C n+1 . The contact<br />

condition for ξ corresponds to the positive definiteness of that Levi form, or what<br />

in complex analysis is called the strict pseudoconvexity of the hypersurface. For<br />

more on the question of pseudoconvexity from the contact geometric viewpoint<br />

see [1, Section 8.2]. Beware that the ‘complex structure’ in their Proposition 8.14<br />

is not required to be integrable, i.e. constitutes what is more commonly referred<br />

to as an ‘almost complex structure’.<br />

Definition 2.11. Let (V, ω) be a symplectic manifold of dimension 2n + 2,<br />

that is, ω is a closed (dω = 0) and non-degenerate (ω n+1 �= 0) 2–form on V . A<br />

vector field X is called a Liouville vector field if LXω = ω, where L denotes<br />

the Lie <strong>der</strong>ivative.<br />

With the help of Cartan’s formula LX = d ◦ iX + iX ◦ d this may be rewrit-<br />

ten as d(iXω) = ω. Then the 1–form α = iXω defines a contact form on any<br />

hypersurface M in V transverse to X. Indeed,<br />

α ∧ (dα) n = iXω ∧ (d(iXω)) n = iXω ∧ ω n = 1<br />

n + 1 iX(ω n+1 ),<br />

which is a volume form on M ⊂ V provided M is transverse to X.<br />

Example 2.12. With V = R 2n+2 , symplectic form ω = �<br />

j dxj ∧ dyj, and<br />

Liouville vector field X = �<br />

j (xj∂xj + yj∂yj )/2 = r∂r/2, we recover the standard<br />

contact structure on S 2n+1 .<br />

For finer issues relating to hypersurfaces in symplectic manifolds transverse<br />

to a Liouville vector field I refer the rea<strong>der</strong> to [1, Section 8.2].<br />

Here is a further useful example of contactomorphic manifolds.<br />

Proposition 2.13. For any point p ∈ S 2n+1 , the manifold (S 2n+1 \ {p}, ξ0) is<br />

contactomorphic to (R 2n+1 , ξ2).<br />

8


Proof. The contact manifold (S 2n+1 , ξ0) is a homogeneous space un<strong>der</strong> the nat-<br />

ural U(n + 1)–action, so we are free to choose p = (0, . . .,0, −1). Stereographic<br />

projection from p does almost, but not quite yield the desired contactomorphism.<br />

Instead, we use a map that is well-known in the theory of Siegel domains (cf. [3,<br />

Chapter 8]) and that looks a bit like a complex analogue of stereographic projec-<br />

tion; this was suggested in [92, Exercise 3.64].<br />

Regard S 2n+1 as the unit sphere in C n+1 = C n ×C with cartesian coordinates<br />

(z1, . . .,zn, w) = (z, w). We identify R 2n+1 with C n ×R ⊂ C n ×C with coordinates<br />

(ζ1, . . .,ζn, s) = (ζ, s) = (ζ,Re σ), where ζj = xj + iyj. Then<br />

n�<br />

α2 = ds + (xj dyj − yj dxj)<br />

and<br />

j=1<br />

= ds + i<br />

(ζ dζ − ζ dζ).<br />

2<br />

α0 = i<br />

(z dz − z dz + w dw − w dw).<br />

2<br />

Now define a smooth map f : S 2n+1 \ {(0, −1)} → R 2n+1 by<br />

Then<br />

and<br />

(ζ, s) = f(z, w) =<br />

�<br />

z i(w − w)<br />

, −<br />

1 + w 2|1 + w| 2<br />

�<br />

.<br />

f ∗ i dw i dw<br />

ds = − +<br />

2|1 + w| 2 2|1 + w| 2<br />

+ i(w − w) dw i(w − w) dw<br />

+<br />

2(1 + w) |1 + w| 2 2(1 + w) |1 + w| 2<br />

i<br />

=<br />

2|1 + w| 2<br />

�<br />

w − w w − w<br />

−dw + dw + dw +<br />

1 + w 1 + w dw<br />

�<br />

f ∗ (ζ dζ − ζ dζ) =<br />

Along S 2n+1 we have<br />

=<br />

�<br />

z dz<br />

1 + w<br />

− z<br />

�<br />

dz<br />

z<br />

(1 + w) 2dw<br />

�<br />

1 + w −<br />

1 + w 1 + w −<br />

z<br />

(1 + w) 2dw<br />

�<br />

1<br />

|1 + w| 2<br />

�<br />

z dz − zdz + |z| 2<br />

� ��<br />

dw dw<br />

− .<br />

1 + w 1 + w<br />

|z| 2 = 1 − |w| 2 = (1 − w)(1 + w) + (w − w)<br />

= (1 − w)(1 + w) − (w − w),<br />

9


whence<br />

|z| 2<br />

� �<br />

dw dw<br />

−<br />

1 + w 1 + w<br />

w − w<br />

= (1 − w)dw −<br />

1 + w dw<br />

w − w<br />

− (1 − w)dw −<br />

1 + w dw.<br />

From these calculations we conclude f ∗ α2 = α0/|1 + w| 2 . So it only remains to<br />

show that f is actually a diffeomorphism of S 2n+1 \ {(0, −1)} onto R 2n+1 . To<br />

that end, consi<strong>der</strong> the map<br />

defined by<br />

�f : (C n × C) \ (C n × {−1}) −→ (C n × C) \ (C n × {−i/2})<br />

(ζ, σ) = � f(z, w) =<br />

� �<br />

z i w − 1<br />

, − .<br />

1 + w 2 w + 1<br />

This is a biholomorphic map with inverse map<br />

� �<br />

2ζ 1 + 2iσ<br />

(ζ, σ) ↦−→ , .<br />

1 − 2iσ 1 − 2iσ<br />

We compute<br />

w − 1 w − 1<br />

Im σ = − −<br />

4(w + 1) 4(w + 1)<br />

(w − 1)(w + 1) + (w − 1)(w + 1)<br />

= −<br />

4|1 + w| 2<br />

= 1 − |w|2<br />

2|1 + w| 2.<br />

Hence for (z, w) ∈ S 2n+1 \ {(0, −1)} we have<br />

Imσ =<br />

|z| 2 1<br />

=<br />

2|1 + w| 2 2 |ζ|2 ;<br />

conversely, any point (ζ, σ) with Imσ = |ζ| 2 /2 lies in the image of � f| S 2n+1 \{(0,−1)},<br />

that is, � f restricted to S 2n+1 \{(0, −1)} is a diffeomorphism onto {Im σ = |ζ| 2 /2}.<br />

Finally, we compute<br />

i(w − 1) i(w − 1)<br />

Re σ = − +<br />

4(w + 1) 4(w + 1)<br />

(w − 1)(w + 1) − (w − 1)(w + 1)<br />

= −i<br />

4|1 + w| 2<br />

i(w − w)<br />

= −<br />

2|1 + w| 2,<br />

10


from which we see that for (z, w) ∈ S 2n+1 \ {(0, −1)} and with (ζ, σ) = � f(z, w)<br />

we have f(z, w) = (ζ,Re σ). This concludes the proof.<br />

At the beginning of this section I mentioned that one may allow contact<br />

structures that are not coorientable, and hence not defined by a global contact<br />

form.<br />

Example 2.14. Let M = R n+1 ×RP n with cartesian coordinates (x0, . . .,xn) on<br />

the R n+1 –factor and homogeneous coordinates [y0 : . . . : yn] on the RP n –factor.<br />

Then<br />

ξ = ker � n �<br />

j=0<br />

�<br />

yj dxj<br />

is a well-defined hyperplane field on M, because the 1–form on the right-hand side<br />

is well-defined up to scaling by a non-zero real constant. On the open submanifold<br />

Uk = {yk �= 0} ∼ = Rn+1 × Rn of M we have ξ = kerαk with<br />

αk = dxk + �<br />

� �<br />

yj<br />

dxj<br />

j�=k<br />

an honest 1–form on Uk. This is the standard contact form of Example 2.6, which<br />

proves that ξ is a contact structure on M.<br />

If n is even, then M is not orientable, so there can be no global contact<br />

form defining ξ (cf. Remark 2.2), i.e. ξ is not coorientable. Notice, however, that<br />

a contact structure on a manifold of dimension 2n + 1 with n even is always<br />

orientable: the sign of (dα) n |ξ does not depend on the choice of local 1–form<br />

defining ξ.<br />

If n is odd, then M is orientable, so it would be possible that ξ is the kernel<br />

of a globally defined 1–form. However, since the sign of α ∧ (dα) n , for n odd, is<br />

independent of the choice of local 1–form defining ξ, it is also conceivable that no<br />

global contact form exists. (In fact, this consi<strong>der</strong>ation shows that any manifold<br />

of dimension 2n + 1, with n odd, admitting a contact structure (coorientable or<br />

not) needs to be orientable.) This is indeed what happens, as we shall prove now.<br />

Proposition 2.15. Let (M, ξ) be the contact manifold of the preceding example.<br />

Then TM/ξ can be identified with the canonical line bundle on RP n (pulled back<br />

to M). In particular, TM/ξ is a non-trivial line bundle, so ξ is not coorientable.<br />

11<br />

yk


Proof. For given y = [y0 : . . . : yn] ∈ RP n , the vector y0∂x0 +· · ·+yn∂xn ∈ TxR n+1<br />

is well-defined up to a non-zero real factor (and independent of x ∈ R n+1 ), and<br />

hence defines a line ℓy in TxR n+1 ∼ = R n+1 . The set<br />

E = {(t, x, y): x ∈ R n+1 , y ∈ RP n , t ∈ ℓy}<br />

⊂ TR n+1 × RP n ⊂ T(R n+1 × RP n ) = TM<br />

with projection (t, x, y) ↦→ (x, y) defines a line sub-bundle of TM that restricts<br />

to the canonical line bundle over {x} × RP n ≡ RP n for each x ∈ R n+1 . The<br />

canonical line bundle over RP n is well-known to be non-trivial [95, p. 16], so the<br />

same holds for E.<br />

Moreover, E is clearly complementary to ξ, i.e. TM/ξ ∼ = E, since<br />

n�<br />

j=0<br />

yj dxj(<br />

n�<br />

k=0<br />

yk∂xk ) =<br />

This proves that that ξ is not coorientable.<br />

n�<br />

j=0<br />

y 2 j �= 0.<br />

To sum up, in the example above we have one of the following two situations:<br />

• If n is odd, then M is orientable; ξ is neither orientable nor coorientable.<br />

• If n is even, then M is not orientable; ξ is not coorientable, but it is ori-<br />

entable.<br />

We close this section with the definition of the most important types of sub-<br />

manifolds.<br />

Definition 2.16. Let (M, ξ) be a contact manifold.<br />

(i) A submanifold L of (M, ξ) is called an isotropic submanifold if TxL ⊂ ξx<br />

for all x ∈ L.<br />

(ii) A submanifold M ′ of M with contact structure ξ ′ is called a contact<br />

submanifold if TM ′ ∩ ξ|M ′ = ξ′ .<br />

Observe that if ξ = kerα and i: M ′ → M denotes the inclusion map, then the<br />

condition for (M ′ , ξ ′ ) to be a contact submanifold of (M, ξ) is that ξ ′ = ker(i ∗ α).<br />

In particular, ξ ′ ⊂ ξ|M ′ is a symplectic sub-bundle with respect to the symplectic<br />

bundle structure on ξ given by dα.<br />

The following is a manifestation of the maximal non-integrability of contact<br />

structures.<br />

12


Proposition 2.17. Let (M, ξ) be a contact manifold of dimension 2n + 1 and L<br />

an isotropic submanifold. Then dim L ≤ n.<br />

Proof. Write i for the inclusion of L in M and let α be an (at least locally<br />

defined) contact form defining ξ. Then the condition for L to be isotropic becomes<br />

i ∗ α ≡ 0. It follows that i ∗ dα ≡ 0. In particular, TpL ⊂ ξp is an isotropic<br />

subspace of the symplectic vector space (ξp, dα|ξp ), i.e. a subspace on which the<br />

symplectic form restricts to zero. From Linear Algebra we know that this implies<br />

dimTpL ≤ (dimξp)/2 = n.<br />

Definition 2.18. An isotropic submanifold L ⊂ (M 2n+1 , ξ) of maximal possible<br />

dimension n is called a Legendrian submanifold.<br />

In particular, in a 3–dimensional contact manifold there are two distinguished<br />

types of knots: Legendrian knots on the one hand, transverse 1 knots on the<br />

other, i.e. knots that are everywhere transverse to the contact structure. If ξ<br />

is cooriented by a contact form α and γ : S 1 → (M, ξ = kerα) is oriented, one<br />

can speak of a positively or negatively transverse knot, depending on whether<br />

α(˙γ) > 0 or α(˙γ) < 0.<br />

2.2 Gray stability and the Moser trick<br />

The Gray stability theorem that we are going to prove in this section says that<br />

there are no non-trivial deformations of contact structures on closed manifolds.<br />

In fancy language, this means that contact structures on closed manifolds have<br />

discrete moduli. First a preparatory lemma.<br />

Lemma 2.19. Let ωt, t ∈ [0, 1], be a smooth family of differential k–forms on a<br />

manifold M and (ψt) t∈[0,1] an isotopy of M. Define a time-dependent vector field<br />

Xt on M by Xt ◦ ψt = ˙ ψt, where the dot denotes <strong>der</strong>ivative with respect to t (so<br />

that ψt is the flow of Xt). Then<br />

d � � � � ∗ ∗<br />

ψt ωt = ψt ˙ωt + LXtωt .<br />

dt<br />

Proof. For a time-independent k–form ω we have<br />

d � ∗<br />

ψt ω<br />

dt<br />

� = ψ ∗� t LXtω � .<br />

This follows by observing that<br />

1 Some people like to call them ‘transversal knots’, but I adhere to J.H.C. Whitehead’s dictum,<br />

as quoted in [64]: “Transversal is a noun; the adjective is transverse.”<br />

13


(i) the formula holds for functions,<br />

(ii) if it holds for differential forms ω and ω ′ , then also for ω ∧ ω ′ ,<br />

(iii) if it holds for ω, then also for dω,<br />

(iv) locally functions and differentials of functions generate the algebra of dif-<br />

ferential forms.<br />

We then compute<br />

d<br />

dt (ψ∗ ψ<br />

t ωt) = lim<br />

h→0<br />

∗ t+hωt+h − ψ∗ t ωt<br />

h<br />

= lim<br />

h→0<br />

ψ ∗ t+h ωt+h − ψ ∗ t+h ωt + ψ ∗ t+h ωt − ψ ∗ t ωt<br />

= lim ψ<br />

h→0 ∗ � �<br />

ωt+h − ωt<br />

t+h + lim<br />

h h→0<br />

� �<br />

˙ωt + LXtωt .<br />

= ψ ∗ t<br />

h<br />

ψ ∗ t+h ωt − ψ ∗ t ωt<br />

h<br />

For that last equality observe (regarding the second summand) that ψt+h =<br />

ψt h ◦ ψt, where ψt h<br />

dependent vector field Xt h := Xt+h; then apply the result for time-independent<br />

k–forms.<br />

denotes, for fixed t and time-variable h, the flow of the time-<br />

Theorem 2.20 (Gray stability). Let ξt, t ∈ [0, 1], be a smooth family of contact<br />

structures on a closed manifold M. Then there is an isotopy (ψt) t∈[0,1] of M such<br />

that<br />

Tψt(ξ0) = ξt for each t ∈ [0, 1].<br />

Proof. The simplest proof of this result rests on what is known as the Moser<br />

trick, introduced by J. Moser [96] in the context of stability results for (equicoho-<br />

mologous) volume and symplectic forms. J. Gray’s original proof [61] was based<br />

on deformation theory à la Kodaira-Spencer. The idea of the Moser trick is to<br />

assume that ψt is the flow of a time-dependent vector field Xt. The desired equa-<br />

tion for ψt then translates into an equation for Xt. If that equation can be solved,<br />

the isotopy ψt is found by integrating Xt; on a closed manifold the flow of Xt will<br />

be globally defined.<br />

Let αt be a smooth family of 1–forms with kerαt = ξt. The equation in the<br />

theorem then translates into<br />

ψ ∗ t αt = λtα0,<br />

14


where λt: M → R + is a suitable smooth family of smooth functions. Differen-<br />

tiation of this equation with respect to t yields, with the help of the preceding<br />

lemma,<br />

ψ ∗� �<br />

t ˙αt + LXtαt = λtα0<br />

˙ = ˙ λt<br />

ψ<br />

λt<br />

∗ t αt,<br />

or, with the help of Cartan’s formula LX = d ◦ iX + iX ◦ d and with µt =<br />

d<br />

dt (log λt) ◦ ψ −1<br />

t ,<br />

ψ ∗� � ∗<br />

t ˙αt + d(αt(Xt)) + iXtdαt = ψt (µtαt).<br />

If we choose Xt ∈ ξt, this equation will be satisfied if<br />

Plugging in the Reeb vector field Rαt gives<br />

˙αt + iXtdαt = µtαt. (2.1)<br />

˙αt(Rαt) = µt. (2.2)<br />

So we can use (2.2) to define µt, and then the non-degeneracy of dαt|ξt and<br />

the fact that Rαt ∈ ker(µtαt − ˙αt) allow us to find a unique solution Xt ∈ ξt<br />

of (2.1).<br />

Remark 2.21. (1) <strong>Contact</strong> forms do not satisfy stability, that is, in general<br />

one cannot find an isotopy ψt such that ψ ∗ t αt = α0. For instance, consi<strong>der</strong> the<br />

following family of contact forms on S 3 ⊂ R 4 :<br />

αt = (x1 dy1 − y1 dx1) + (1 + t)(x2 dy2 − y2 dx2),<br />

where t ≥ 0 is a real parameter. The Reeb vector field of αt is<br />

Rαt = (x1 ∂y1 − y1<br />

1<br />

∂x1 ) +<br />

1 + t (x2 ∂y2 − y2 ∂x2 ).<br />

The flow of Rα0 defines the Hopf fibration, in particular all orbits of Rα0 are<br />

closed. For t ∈ R + \ Q, on the other hand, Rαt has only two periodic orbits. So<br />

there can be no isotopy with ψ ∗ t αt = α0, because such a ψt would also map Rα0<br />

to Rαt.<br />

(2) Y. Eliashberg [25] has shown that on the open manifold R 3 there are<br />

likewise no non-trivial deformations of contact structures, but on S 1 × R 2 there<br />

does exist a continuum of non-equivalent contact structures.<br />

(3) For further applications of this theorem it is useful to observe that at<br />

points p ∈ M with ˙αt,p identically zero in t we have Xt(p) ≡ 0, so such points<br />

remain stationary un<strong>der</strong> the isotopy ψt.<br />

15


2.3 <strong>Contact</strong> Hamiltonians<br />

A vector field X on the contact manifold (M, ξ = kerα) is called an infinitesi-<br />

mal automorphism of the contact structure if the local flow of X preserves ξ<br />

(The study of such automorphisms was initiated by P. Libermann, cf. [80]). By<br />

slight abuse of notation, we denote this flow by ψt; if M is not closed, ψt (for a<br />

fixed t �= 0) will not in general be defined on all of M. The condition for X to<br />

be an infinitesimal automorphism can be written as Tψt(ξ) = ξ, which is equiv-<br />

alent to LXα = λα for some function λ: M → R (notice that this condition is<br />

independent of the choice of 1–form α defining ξ). The local flow of X preserves<br />

α if and only if LXα = 0.<br />

Theorem 2.22. With a fixed choice of contact form α there is a one-to-one<br />

correspondence between infinitesimal automorphisms X of ξ = kerα and smooth<br />

functions H : M → R. The correspondence is given by<br />

• X ↦−→ HX = α(X);<br />

• H ↦−→ XH, defined uniqely by α(XH) = H and iXH dα = dH(Rα)α − dH.<br />

The fact that XH is uniquely defined by the equations in the theorem follows<br />

as in the preceding section from the fact that dα is non-degenerate on ξ and<br />

Rα ∈ ker(dH(Rα)α − dH).<br />

Proof. Let X be an infinitesimal automorphism of ξ. Set HX = α(X) and write<br />

dHX + iXdα = LXα = λα with λ: M → R. Applying this last equation to<br />

Rα yields dHX(Rα) = λ. So X satisfies the equations α(X) = HX and iXdα =<br />

dHX(Rα)α − dHX. This means that XHX = X.<br />

Conversely, given H : M → R and with XH as defined in the theorem, we<br />

have<br />

LXH α = iXH dα + d(α(XH)) = dH(Rα)α,<br />

so XH is an infinitesimal automorphism of ξ. Moreover, it is immediate from the<br />

definitions that HXH = α(XH) = H.<br />

Corollary 2.23. Let (M, ξ = kerα) be a closed contact manifold and Ht: M →<br />

R, t ∈ [0, 1], a smooth family of functions. Let Xt = XHt be the correspond-<br />

ing family of infinitesimal automorphisms of ξ (defined via the correspondence<br />

described in the preceding theorem). Then the globally defined flow ψt of the<br />

16


time-dependent vector field Xt is a contact isotopy of (M, ξ), that is, ψ ∗ t α = λtα<br />

for some smooth family of functions λt: M → R + .<br />

Proof. With Lemma 2.19 and the preceding proof we have<br />

d � ∗<br />

ψt α<br />

dt<br />

� = ψ ∗� t LXtα � = ψ ∗� t dHt(Rα)α � = µtψ ∗ t α<br />

with µt = dHt(Rα) ◦ ψt. Since ψ0 = idM (whence ψ∗ 0α = α) this implies that,<br />

with<br />

we have ψ ∗ t α = λtα.<br />

λt = exp �� t<br />

µs ds � ,<br />

0<br />

This corollary will be used in Section 2.5 to prove various isotopy extension<br />

theorems from isotopies of special submanifolds to isotopies of the ambient con-<br />

tact manifold. In a similar vein, contact Hamiltonians can be used to show that<br />

standard general position arguments from differential topology continue to hold<br />

in the contact geometric setting. Another application of contact Hamiltonians<br />

is a proof of the fact that the contactomorphism group of a connected contact<br />

manifold acts transitively on that manifold [12]. (See [8] for more on the general<br />

structure of contactomorphism groups.)<br />

2.4 Darboux’s theorem and neighbourhood theorems<br />

The flexibility of contact structures inherent in the Gray stability theorem and<br />

the possibility to construct contact isotopies via contact Hamiltonians results in<br />

a variety of theorems that can be summed up as saying that there are no local<br />

invariants in contact geometry. Such theorems form the theme of the present<br />

section.<br />

In contrast with Riemannian geometry, for instance, where the local structure<br />

coming from the curvature gives rise to a rich theory, the interesting questions<br />

in contact geometry thus appear only at the global level. However, it is actually<br />

that local flexibility that allows us to prove strong global theorems, such as the<br />

existence of contact structures on certain closed manifolds.<br />

2.4.1 Darboux’s theorem<br />

Theorem 2.24 (Darboux’s theorem). Let α be a contact form on the (2n +<br />

1)–dimensional manifold M and p a point on M. Then there are coordinates<br />

17


x1, . . .,xn, y1, . . .,yn, z on a neighbourhood U ⊂ M of p such that<br />

α|U = dz +<br />

n�<br />

j=1<br />

xj dyj.<br />

Proof. We may assume without loss of generality that M = R 2n+1 and p = 0 is<br />

the origin of R 2n+1 . Choose linear coordinates x1, . . .,xn, y1, . . .yn, z on R 2n+1<br />

such that<br />

on T0R 2n+1 :<br />

�<br />

α(∂z) = 1, i∂zdα = 0,<br />

∂xj , ∂yj ∈ kerα (j = 1, . . .,n), dα = � n<br />

j=1 dxj ∧ dyj.<br />

This is simply a matter of linear algebra (the normal form theorem for skew-<br />

symmetric forms on a vector space).<br />

Now set α0 = dz + �<br />

j xj dyj and consi<strong>der</strong> the family of 1–forms<br />

αt = (1 − t)α0 + tα, t ∈ [0, 1],<br />

on R 2n+1 . Our choice of coordinates ensures that<br />

αt = α, dαt = dα at the origin.<br />

Hence, on a sufficiently small neighbourhood of the origin, αt is a contact form<br />

for all t ∈ [0, 1].<br />

We now want to use the Moser trick to find an isotopy ψt of a neighbourhood<br />

of the origin such that ψ ∗ t αt = α0. This aim seems to be in conflict with our<br />

earlier remark that contact forms are not stable, but as we shall see presently,<br />

locally this equation can always be solved.<br />

Indeed, differentiating ψ ∗ t αt = α0 (and assuming that ψt is the flow of some<br />

time-dependent vector field Xt) we find<br />

so Xt needs to satisfy<br />

ψ ∗� �<br />

t ˙αt + LXtαt = 0,<br />

˙αt + d(αt(Xt)) + iXtdαt = 0. (2.3)<br />

Write Xt = HtRαt + Yt with Yt ∈ ker αt. Inserting Rαt in (2.3) gives<br />

˙αt(Rαt) + dHt(Rαt) = 0. (2.4)<br />

18


On a neighbourhood of the origin, a smooth family of functions Ht satisfying<br />

(2.4) can always be found by integration, provided only that this neighbourhood<br />

has been chosen so small that none of the Rαt has any closed orbits there. Since<br />

αt ˙ is zero at the origin, we may require that Ht(0) = 0 and dHt|0 = 0 for all<br />

t ∈ [0, 1]. Once Ht has been chosen, Yt is defined uniquely by (2.3), i.e. by<br />

˙αt + dHt + iYtdαt = 0.<br />

Notice that with our assumptions on Ht we have Xt(0) = 0 for all t.<br />

Now define ψt to be the local flow of Xt. This local flow fixes the origin, so<br />

there it is defined for all t ∈ [0, 1]. Since the domain of definition in R × M of a<br />

local flow on a manifold M is always open (cf. [15, 8.11]), we can infer 2 that ψt<br />

is actually defined for all t ∈ [0, 1] on a sufficiently small neighbourhood of the<br />

origin in R 2n+1 . This concludes the proof of the theorem (strictly speaking, the<br />

local coordinates in the statement of the theorem are the coordinates xj ◦ ψ −1<br />

1<br />

etc.).<br />

Remark 2.25. The proof of this result given in [1] is incomplete: It is not<br />

possible, as is suggested there, to prove the Darboux theorem for contact forms<br />

if one requires Xt ∈ ker αt.<br />

2.4.2 Isotropic submanifolds<br />

Let L ⊂ (M, ξ = kerα) be an isotropic submanifold in a contact manifold with<br />

cooriented contact structure. Write (TL) ⊥ ⊂ ξ|L for the sub-bundle of ξ|L that is<br />

symplectically orthogonal to TL with respect to the symplectic bundle structure<br />

dα|ξ. The conformal class of this symplectic bundle structure depends only on<br />

the contact structure ξ, not on the choice of contact form α defining ξ: If α is<br />

replaced by λα for some smooth function λ: M → R + , then d(λα)|ξ = λ dα|ξ.<br />

So the bundle (TL) ⊥ is determined by ξ.<br />

The fact that L is isotropic implies TL ⊂ (TL) ⊥ . Following Weinstein [105],<br />

we call the quotient bundle (TL) ⊥ /TL with the conformal symplectic structure<br />

induced by dα the conformal symplectic normal bundle of L in M and write<br />

CSN(M, L) = (TL) ⊥ /TL.<br />

2 To be absolutely precise, one ought to work with a family αt, t ∈ R, where αt ≡ α0 for<br />

t ≤ ε and αt ≡ α1 for t ≥ 1 − ε, i.e. a technical homotopy in the sense of [15]. Then Xt will be<br />

defined for all t ∈ R, and the reasoning of [15] can be applied.<br />

19


So the normal bundle NL = (TM|L)/TL of L in M can be split as<br />

NL ∼ = (TM|L)/(ξ|L) ⊕ (ξ|L)/(TL) ⊥ ⊕ CSN(M, L).<br />

Observe that if dimM = 2n + 1 and dimL = k ≤ n, then the ranks of the<br />

three summands in this splitting are 1, k and 2(n − k), respectively. Our aim<br />

in this section is to show that a neighbourhood of L in M is determined, up to<br />

contactomorphism, by the isomorphism type (as a conformal symplectic bundle)<br />

of CSN(M, L).<br />

The bundle (TM|L)/(ξ|L) is a trivial line bundle because ξ is cooriented.<br />

The bundle (ξ|L)/(TL) ⊥ can be identified with the cotangent bundle T ∗ L via the<br />

well-defined bundle isomorphism<br />

Ψ: (ξ|L)/(TL) ⊥ −→ T ∗ L<br />

Y ↦−→ iY dα|TL.<br />

(Ψ is obviously injective and well-defined by the definition of (TL) ⊥ , and the<br />

ranks of the two bundles are equal.)<br />

Although Ψ is well-defined on the quotient (ξ|L)/(TL) ⊥ , to proceed further<br />

we need to choose an isotropic complement of (TL) ⊥ in ξ|L. Restricted to each<br />

fibre ξp, p ∈ L, such an isotropic complement of (TpL) ⊥ exists. There are two<br />

ways to obtain a smooth bundle of such isotropic complements. The first would<br />

be to carry over Arnold’s corresponding discussion of Lagrangian subbundles<br />

of symplectic bundles [6] to the isotropic case in or<strong>der</strong> to show that the space<br />

of isotropic complements of U ⊥ ⊂ V , where U is an isotropic subspace in a<br />

symplectic vector space V , is convex. (This argument uses generating functions<br />

for isotropic subspaces.) Then by a partition of unity argument the desired<br />

complement can be constructed on the bundle level.<br />

A slightly more pedestrian approach is to define this isotropic complement<br />

with the help of a complex bundle structure J on ξ compatible with dα (cf.<br />

Remark 2.3). The condition dα(X, JX) > 0 for 0 �= X ∈ ξ implies that (TpL) ⊥ ∩<br />

J(TpL) = {0} for all p ∈ L, and so a dimension count shows that J(TL) is indeed<br />

a complement of (TL) ⊥ in ξ|L. (In a similar vein, CSN(M, L) can be identified<br />

as a sub-bundle of ξ, viz., the orthogonal complement of TL ⊕ J(TL) ⊂ ξ with<br />

respect to the bundle metric dα(., J.) on ξ.)<br />

On the Whitney sum TL ⊕ T ∗ L (for any manifold L) there is a canonical<br />

symplectic bundle structure ΩL defined by<br />

ΩL,p(X + η, X ′ + η ′ ) = η(X ′ ) − η ′ (X) for X, X ′ ∈ TpL; η, η ′ ∈ T ∗ p L.<br />

20


Lemma 2.26. The bundle map<br />

idTL ⊕ Ψ: (TL ⊕ J(TL), dα) −→ (TL ⊕ T ∗ L,ΩL)<br />

is an isomorphism of symplectic vector bundles.<br />

Proof. We only need to check that idTL ⊕ Ψ is a symplectic bundle map. Let<br />

X, X ′ ∈ TpL and Y, Y ′ ∈ Jp(TpL). Write Y = JpZ, Y ′ = JpZ ′ with Z, Z ′ ∈ TpL.<br />

It follows that<br />

dα(Y, Y ′ ) = dα(JZ, JZ ′ ) = dα(Z, Z ′ ) = 0,<br />

since L is an isotropic submanifold. For the same reason dα(X, X ′ ) = 0. Hence<br />

dα(X + Y, X ′ + Y ′ ) = dα(Y, X ′ ) − dα(Y ′ , X)<br />

= Ψ(Y )(X ′ ) − Ψ(Y ′ )(X)<br />

= ΩL(X + Ψ(Y ), X ′ + Ψ(Y ′ )).<br />

Theorem 2.27. Let (Mi, ξi), i = 0, 1, be contact manifolds with closed isotropic<br />

submanifolds Li. Suppose there is an isomorphism of conformal symplectic nor-<br />

mal bundles Φ: CSN(M0, L0) → CSN(M1, L1) that covers a diffeomorphism<br />

φ: L0 → L1. Then φ extends to a contactomorphism ψ: N(L0) → N(L1) of<br />

suitable neighbourhoods N(Li) of Li such that Tψ| CSN(M0,L0) and Φ are bundle<br />

homotopic (as symplectic bundle isomorphisms).<br />

Corollary 2.28. Diffeomorphic (closed) Legendrian submanifolds have contac-<br />

tomorphic neighbourhoods.<br />

Proof. If Li ⊂ Mi is Legendrian, then CSN(Mi, Li) has rank 0, so the conditions<br />

in the theorem, apart from the existence of a diffeomorphism φ: L0 → L1, are<br />

void.<br />

Example 2.29. Let S 1 ⊂ (M 3 , ξ) be a Legendrian knot in a contact 3–manifold.<br />

Then with a coordinate θ ∈ [0, 2π] along S 1 and coordinates x, y in slices trans-<br />

verse to S 1 , the contact structure<br />

cos θ dx − sin θ dy = 0<br />

provides a model for a neighbourhood of S 1 .<br />

21


Proof of Theorem 2.27. Choose contact forms αi for ξi, i = 0, 1, scaled in such a<br />

way that Φ is actually an isomorphism of symplectic vector bundles with respect<br />

to the symplectic bundle structures on CSN(Mi, Li) given by dαi. Here we think<br />

of CSN(Mi, Li) as a sub-bundle of TMi|Li (rather than as a quotient bundle).<br />

We identify (TMi|Li )/(ξi|Li ) with the trivial line bundle spanned by the Reeb<br />

vector field Rαi . In total, this identifies<br />

as a sub-bundle of TMi|Li .<br />

NLi = 〈Rαi 〉 ⊕ Ji(TLi) ⊕ CSN(Mi, Li)<br />

Let ΦR: 〈Rα0 〉 → 〈Rα1 〉 be the obvious bundle isomorphism defined by requiring<br />

that Rα0 (p) map to Rα1 (φ(p)).<br />

Let Ψi: Ji(TLi) → T ∗Li be the isomorphism defined by taking the interior<br />

product with dαi. Notice that<br />

Tφ ⊕ (φ ∗ ) −1 : (TL0 ⊕ T ∗ L0, ΩL0 ) → (TL1 ⊕ T ∗ L1, ΩL1 )<br />

is an isomorphism of symplectic vector bundles. With Lemma 2.26 it follows that<br />

Tφ ⊕ Ψ −1<br />

1 ◦ (φ∗ ) −1 ◦ Ψ0: (TL0 ⊕ J0(TL0), dα0) → (TL1 ⊕ J1(TL1), dα1)<br />

is an isomorphism of symplectic vector bundles.<br />

Now let<br />

�Φ: NL0 −→ NL1<br />

be the bundle isomorphism (covering φ) defined by<br />

�Φ = ΦR ⊕ Ψ −1<br />

1 ◦ (φ∗ ) −1 ◦ Ψ0 ⊕ Φ.<br />

Let τi: NLi → Mi be tubular maps, that is, the τ (I suppress the index i for<br />

better readability) are embeddings such that τ|L – where L is identified with the<br />

zero section of NL – is the inclusion L ⊂ M, and Tτ induces the identity on NL<br />

along L (with respect to the splittings T(NL)|L = TL ⊕ NL = TM|L).<br />

Then τ1 ◦ � Φ ◦ τ −1<br />

0 : N(L0) → N(L1) is a diffeomorphism of suitable neighbourhoods<br />

N(Li) of Li that induces the bundle map<br />

Tφ ⊕ � Φ: TM0|L0 −→ TM1|L1 .<br />

By construction, this bundle map pulls α1 back to α0 and dα1 to dα0. Hence, α0<br />

and (τ1 ◦ � Φ ◦ τ −1<br />

0 )∗α1 are contact forms on N(L0) that coincide on TM0|L0 , and<br />

so do their differentials.<br />

22


Now consi<strong>der</strong> the family of 1–forms<br />

βt = (1 − t)α0 + t(τ1 ◦ � Φ ◦ τ −1<br />

0 )∗ α1, t ∈ [0, 1].<br />

On TM0|L0 we have βt ≡ α0 and dβt ≡ dα0. Since the contact condition α ∧<br />

(dα) n �= 0 is an open condition, we may assume – shrinking N(L0) if necessary<br />

– that βt is a contact form on N(L0) for all t ∈ [0, 1]. By the Gray stability<br />

theorem (Thm. 2.20) and Remark 2.21 (3) following its proof, we find an isotopy<br />

ψt of N(L0), fixing L0, such that ψ ∗ t βt = λtα0 for some smooth family of smooth<br />

functions λt: N(L0) → R + .<br />

(Since N(L0) is not a closed manifold, ψt is a priori only a local flow. But<br />

on L0 it is stationary and hence defined for all t. As in the proof of the Darboux<br />

theorem (Thm. 2.24) we conclude that ψt is defined for all t ∈ [0, 1] in a sufficiently<br />

small neighbourhood of L0, so shrinking N(L0) once again, if necessary, will<br />

ensure that ψt is a global flow on N(L0).)<br />

We conclude that ψ = τ1 ◦ � Φ ◦ τ −1<br />

0 ◦ ψ1 is the desired contactomorphism.<br />

Remark 2.30. With a little more care one can actually achieve Tψ1 = id on<br />

TM0|L0 , which implies in particular that Tψ| CSN(M0,L0) = Φ, cf. [105]. (Remem-<br />

ber that there is a certain freedom in constructing an isotopy via the Moser trick<br />

if the condition Xt ∈ ξt is dropped.) The key point is the generalised Poincaré<br />

lemma, cf. [80, p. 361], which allows us to write a closed differential form γ given<br />

in a neighbourhood of the zero section of a bundle and vanishing along that zero<br />

section as an exact form γ = dη with η and its partial <strong>der</strong>ivatives with respect<br />

to all coordinates (in any chart) vanishing along the zero section. This lemma is<br />

applied first to γ = d(β1 − β0), in or<strong>der</strong> to find (with the symplectic Moser trick)<br />

a diffeomorphism σ of a neighbourhood of L0 ⊂ M0 with Tσ = id on TM0|L0<br />

and such that dβ0 = d(σ ∗ β1). It is then applied once again to γ = β0 − σ ∗ β1.<br />

(The proof of the symplectic neighbourhood theorem in [92] appears to be<br />

incomplete in this respect.)<br />

Example 2.31. Let M0 = M1 = R 3 with contact forms α0 = dz + x dy and<br />

α1 = dz + (x + y)dy and L0 = L1 = 0 the origin in R 3 . Thus<br />

We take Φ = idCSN.<br />

CSN(M0, L0) = CSN(M1, L1) = span{∂x, ∂y} ⊂ T0R 3 .<br />

23


Set αt = dz+(x+ty)dy. The Moser trick with Xt ∈ ker αt yields Xt = −y∂x,<br />

and hence ψt(x, y, z) = (x − ty, y, z). Then<br />

Tψ1 =<br />

⎛<br />

⎜<br />

⎝<br />

which does not restrict to Φ on CSN.<br />

1 −1 0<br />

0 1 0<br />

0 0 1<br />

⎞<br />

⎟<br />

⎠,<br />

However, a different solution for ψ ∗ t αt = α0 is ψt(x, y, z) = (x, y, z − ty 2 /2),<br />

found by integrating Xt = −y 2 ∂z/2 (a multiple of the Reeb vector field of αt).<br />

Here we get<br />

Tψ1 =<br />

⎛<br />

⎜<br />

⎝<br />

1 0 0<br />

0 1 0<br />

0 −y 1<br />

⎞<br />

⎟<br />

⎠ ,<br />

hence Tψ1| T0R 3 = id, so in particular Tψ1|CSN = Φ.<br />

2.4.3 <strong>Contact</strong> submanifolds<br />

Let (M ′ , ξ ′ = kerα ′ ) ⊂ (M, ξ = kerα) be a contact submanifold, that is, TM ′ ∩<br />

ξ|M ′ = ξ′ . As before we write (ξ ′ ) ⊥ ⊂ ξ|M ′ for the symplectically orthogonal<br />

complement of ξ ′ in ξ|M ′. Since M ′ is a contact submanifold (so ξ ′ is a symplectic<br />

sub-bundle of (ξ|M ′, dα)), we have<br />

TM ′ ⊕ (ξ ′ ) ⊥ = TM|M ′,<br />

i.e. we can identify (ξ ′ ) ⊥ with the normal bundle NM ′ . Moreover, dα induces a<br />

conformal symplectic structure on (ξ ′ ) ⊥ , so we call (ξ ′ ) ⊥ the conformal sym-<br />

plectic normal bundle of M ′ in M and write<br />

CSN(M, M ′ ) = (ξ ′ ) ⊥ .<br />

Theorem 2.32. Let (Mi, ξi), i = 0, 1, be contact manifolds with compact contact<br />

submanifolds (M ′ i , ξ′ i ). Suppose there is an isomorphism of conformal symplectic<br />

normal bundles Φ: CSN(M0, M ′ 0 ) → CSN(M1, M ′ 1 ) that covers a contactomorphism<br />

φ: (M ′ 0 , ξ′ 0 ) → (M ′ 1 , ξ′ 1 ). Then φ extends to a contactomorphism ψ of<br />

suitable neighbourhoods N(M ′ i ) of M ′ i such that Tψ| CSN(M0,M ′ 0 ) and Φ are bundle<br />

homotopic (as symplectic bundle isomorphisms) up to a conformality.<br />

24


Example 2.33. A particular instance of this theorem is the case of a transverse<br />

knot in a contact manifold (M, ξ), i.e. an embedding S 1 ֒→ (M, ξ) transverse to ξ.<br />

Since the symplectic group Sp(2n) of linear transformations of R 2n preserving the<br />

standard symplectic structure ω0 = � n<br />

i=1 dxi ∧dyi is connected, there is only one<br />

conformal symplectic R 2n –bundle over S 1 up to conformal equivalence. A model<br />

for the neighbourhood of a transverse knot is given by<br />

� S 1 × R 2n , ξ = ker � dθ +<br />

n�<br />

(xi dyi − yi dxi) �� ,<br />

where θ denotes the S 1 –coordinate; the theorem says that in suitable local coor-<br />

dinates the neighbourhood of any transverse knot looks like this model.<br />

Proof of Theorem 2.32. As in the proof of Theorem 2.27 it is sufficient to find<br />

contact forms αi on Mi and a bundle map TM0| M ′ 0 → TM1| M ′ 1 , covering φ and<br />

inducing Φ, that pulls back α1 to α0 and dα1 to dα0; the proof then concludes<br />

as there with a stability argument.<br />

i=1<br />

For this we need to make a judicious choice of αi. The essential choice is made<br />

separately on each Mi, so I suppress the subscript i for the time being. Choose a<br />

contact form α ′ for ξ ′ on M ′ . Write R ′ for the Reeb vector field of α ′ . Given any<br />

contact form α for ξ on M we may first scale it such that α(R ′ ) ≡ 1 along M ′ .<br />

Then α|TM ′ = α′ , and hence dα|TM ′ = dα′ . We now want to scale α further<br />

such that its Reeb vector field R coincides with R ′ along M ′ . To this end it is<br />

sufficient to find a smooth function f : M → R + with f|M ′ ≡ 1 and iR ′d(fα) ≡ 0<br />

on TM|M ′. This last equation becomes<br />

0 = iR ′d(fα) = iR ′(df ∧ α + f dα) = −df + iR ′dα on TM|M ′.<br />

Since iR ′dα|TM ′ = iR ′dα′ ≡ 0, such an f can be found.<br />

The choices of α ′ 0 and α′ 1 cannot be made independently of each other; we may<br />

first choose α ′ 1 , say, and then define α′ 0 = φ∗α ′ 1 . Then define α0, α1 as described<br />

and scale Φ such that it is a symplectic bundle isomorphism of<br />

Then<br />

((ξ ′ 0) ⊥ , dα0) −→ ((ξ ′ 1) ⊥ , dα1).<br />

Tφ ⊕ Φ: TM0| M ′ 0 −→ TM1| M ′ 1<br />

is the desired bundle map that pulls back α1 to α0 and dα1 to dα0.<br />

25


Remark 2.34. The condition that Ri ≡ R ′ i along M ′ is necessary for ensuring<br />

that (Tφ ⊕ Φ)(R0) = R1, which guarantees (with the other stated conditions)<br />

that (Tφ ⊕ Φ) ∗ (dα1) = dα0. The condition dαi| TM ′ i = dα ′ i<br />

choice of Φ alone would only give (Tφ ⊕ Φ) ∗ (dα1|ξ1 ) = dα0|ξ0 .<br />

2.4.4 Hypersurfaces<br />

and the described<br />

Let S be an oriented hypersurface in a contact manifold (M, ξ = kerα) of dimen-<br />

sion 2n + 1. In a neighbourhood of S in M, which we can identify with S × R<br />

(and S with S × {0}), the contact form α can be written as<br />

α = βr + ur dr,<br />

where βr, r ∈ R, is a smooth family of 1–forms on S and ur : S → R a smooth<br />

family of functions. The contact condition α ∧ (dα) n �= 0 then becomes<br />

0 �= α ∧ (dα) n = (βr + ur dr) ∧ (dβr − ˙ βr ∧ dr + dur ∧ dr) n<br />

= (−nβr ∧ ˙ βr + nβr ∧ dur + ur dβr) ∧ (dβr) n−1 ∧ dr, (2.5)<br />

where the dot denotes <strong>der</strong>ivative with respect to r. The intersection TS ∩ (ξ|S)<br />

determines a distribution (of non-constant rank) of subspaces of TS. If α is<br />

written as above, this distribution is given by the kernel of β0, and hence, at<br />

a given p ∈ S, defines either the full tangent space TpS (if β0,p = 0) or a 1–<br />

codimensional subspace both of TpS and ξp (if β0,p �= 0). In the former case, the<br />

symplectically orthogonal complement (TpS ∩ξp) ⊥ (with respect to the conformal<br />

symplectic structure dα on ξp) is {0}; in the latter case, (TpS ∩ ξp) ⊥ is a 1–<br />

dimensional subspace of ξp contained in TpS ∩ ξp.<br />

From that it is intuitively clear what one should mean by a ‘singular 1–<br />

dimensional foliation’, and we make the following somewhat provisional defini-<br />

tion.<br />

Definition 2.35. The characteristic foliation Sξ of a hypersurface S in (M, ξ)<br />

is the singular 1–dimensional foliation of S defined by (TS ∩ ξ|S) ⊥ .<br />

Example 2.36. If dimM = 3 and dimS = 2, then (TpS ∩ξp) ⊥ = TpS ∩ξp at the<br />

points p ∈ S where TpS ∩ ξp is 1–dimensional. Figure 2 shows the characteristic<br />

foliation of the unit 2–sphere in (R 3 , ξ2), where ξ2 denotes the standard contact<br />

26


structure of Example 2.7: The only singular points are (0, 0, ±1); away from these<br />

points the characteristic foliation is spanned by<br />

(y − xz)∂x − (x + yz)∂y + (x 2 + y 2 )∂z.<br />

Figure 2: The characteristic foliation on S 2 ⊂ (R 3 , ξ2).<br />

The following lemma helps to clarify the notion of singular 1–dimensional<br />

foliation.<br />

Lemma 2.37. Let β0 be the 1–form induced on S by a contact form α defining ξ,<br />

and let Ω be a volume form on S. Then Sξ is defined by the vector field X<br />

satisfying<br />

iXΩ = β0 ∧ (dβ0) n−1 .<br />

Proof. First of all, we observe that β0 ∧ (dβ0) n−1 �= 0 outside the zeros of β0:<br />

Arguing by contradiction, assume β0,p �= 0 and β0 ∧(dβ0) n−1 |p = 0 at some p ∈ S.<br />

Then (dβ0) n |p �= 0 by (2.5). On the codimension 1 subspace kerβ0,p of TpS the<br />

symplectic form dβ0,p has maximal rank n−1. It follows that β0 ∧(dβ0) n−1 |p �= 0<br />

after all, a contradiction.<br />

Next we want to show that X ∈ ker β0. We observe<br />

0 = iX(iXΩ) = β0(X)(dβ0) n−1 − (n − 1)β0 ∧ iXdβ0 ∧ (dβ0) n−2 . (2.6)<br />

Taking the exterior product of this equation with β0 we get<br />

β0(X)β0 ∧ (dβ0) n−1 = 0.<br />

By our previous consi<strong>der</strong>ation this implies β0(X) = 0.<br />

It remains to show that for β0,p �= 0 we have<br />

dβ0(X(p), v) = 0 for all v ∈ TpS ∩ ξp.<br />

27


For n = 1 this is trivially satisfied, because in that case v is a multiple of X(p).<br />

I suppress the point p in the following calculation, where we assume n ≥ 2.<br />

From (2.6) and with β0(X) = 0 we have<br />

Taking the interior product with v ∈ TS ∩ ξ yields<br />

β0 ∧ iXdβ0 ∧ (dβ0) n−2 = 0. (2.7)<br />

−dβ0(X, v)β0 ∧ (dβ0) n−2 + (n − 2)β0 ∧ iXdβ0 ∧ ivdβ0 ∧ (dβ0) n−3 = 0.<br />

(Thanks to the coefficient n − 2 the term (dβ0) n−3 is not a problem for n = 2.)<br />

Taking the exterior product of that last equation with dβ0, and using (2.7), we<br />

find<br />

and thus dβ0(X, v) = 0.<br />

dβ0(X, v)β0 ∧ (dβ0) n−1 = 0,<br />

Remark 2.38. (1) We can now give a more formal definition of ‘singular 1–<br />

dimensional foliation’ as an equivalence class of vector fields [X], where X is<br />

allowed to have zeros and [X] = [X ′ ] if there is a nowhere zero function on all<br />

of S such that X ′ = fX. Notice that the non-integrability of contact structures<br />

and the reasoning at the beginning of the proof of the lemma imply that the zero<br />

set of X does not contain any open subsets of S.<br />

(2) If the contact structure ξ is cooriented rather than just coorientable, so<br />

that α is well-defined up to multiplication with a positive function, then this<br />

lemma allows to give an orientation to the characteristic foliation: Changing α<br />

to λα with λ: M → R + will change β0 ∧ (dβ0) n−1 by a factor λ n .<br />

We now restrict attention to surfaces in contact 3–manifolds, where the notion<br />

of characteristic foliation has proved to be particularly useful.<br />

The following theorem is due to E. Giroux [52].<br />

Theorem 2.39 (Giroux). Let Si be closed surfaces in contact 3–manifolds (Mi, ξi),<br />

i = 0, 1 (with ξi coorientable), and φ: S0 → S1 a diffeomorphism with φ(S0,ξ0 ) =<br />

as oriented characteristic foliations. Then there is a contactomorphism<br />

S1,ξ1<br />

ψ: N(S0) → N(S1) of suitable neighbourhoods N(Si) of Si with ψ(S0) = S1<br />

and such that ψ|S0 is isotopic to φ via an isotopy preserving the characteristic<br />

foliation.<br />

28


Proof. By passing to a double cover, if necessary, we may assume that the Si<br />

are orientable hypersurfaces. Let αi be contact forms defining ξi. Extend φ to a<br />

diffeomorphism (still denoted φ) of neighbourhoods of Si and consi<strong>der</strong> the contact<br />

forms α0 and φ ∗ α1 on a neighbourhood of S0, which we may identify with S0 ×R.<br />

By rescaling α1 we may assume that α0 and φ ∗ α1 induce the same form β0<br />

on S0 ≡ S0 × {0}, and hence also the same form dβ0.<br />

Observe that the expression on the right-hand side of equation (2.5) is linear in<br />

˙βr and ur. This implies that convex linear combinations of solutions of (2.5) (for<br />

n = 1) with the same β0 (and dβ0) will again be solutions of (2.5) for sufficiently<br />

small r. This reasoning applies to<br />

αt := (1 − t)α0 + tφ ∗ α1, t ∈ [0, 1].<br />

(I hope the rea<strong>der</strong> will forgive the slight abuse of notation, with α1 denoting both<br />

a form on M1 and its pull-back φ ∗ α1 to M0.) As is to be expected, we now use<br />

the Moser trick to find an isotopy ψt with ψ ∗ t αt = λtα0, just as in the proof of<br />

Gray stability (Theorem 2.20). In particular, we require as there that the vector<br />

field Xt that we want to integrate to the flow ψt lie in the kernel of αt.<br />

On TS0 we have ˙αt ≡ 0 (thanks to the assumption that α0 and φ ∗ α1 induce<br />

the same form β0 on S0). In particular, if v is a vector in S0,ξ0 , then by equation<br />

(2.1) we have dαt(Xt, v) = 0, which implies that Xt is a multiple of v, hence<br />

tangent to S0,ξ0 . This shows that the flow of Xt preserves S0 and its characteristic<br />

foliation. More formally, we have<br />

LXtαt = d(αt(Xt)) + iXtdαt = iXtdαt,<br />

so with v as above we have LXtαt(v) = 0, which shows that LXtαt|TS0 is a<br />

multiple of α0|TS0 = β0. This implies that the (local) flow of Xt changes β0 by a<br />

conformal factor.<br />

Since S0 is closed, the local flow of Xt restricted to S0 integrates up to t = 1,<br />

and so the same is true 3 in a neighbourhood of S0. Then ψ = φ ◦ ψ1 will be the<br />

desired diffeomorphism N(S0) → N(S1).<br />

As observed previously in the proof of Darboux’s theorem for contact forms,<br />

the Moser trick allows more flexibility if we drop the condition αt(Xt) = 0.<br />

We are now going to exploit this extra freedom to strengthen Giroux’s theorem<br />

3 Cf. the proof (and the footnote therein) of Darboux’s theorem (Thm. 2.24).<br />

29


slightly. This will be important later on when we want to extend isotopies of<br />

hypersurfaces.<br />

Theorem 2.40. Un<strong>der</strong> the assumptions of the preceding theorem we can find<br />

ψ: N(S0) → N(S1) satisfying the stronger condition that ψ|S0 = φ.<br />

Proof. We want to show that the isotopy ψt of the preceding proof may be as-<br />

sumed to fix S0 pointwise. As there, we may assume ˙αt|TS0<br />

≡ 0.<br />

If the condition that Xt be tangent to kerαt is dropped, the condition Xt<br />

needs to satisfy so that its flow will pull back αt to λtα0 is<br />

˙αt + d(αt(Xt)) + iXtdαt = µtαt, (2.8)<br />

where µt and λt are related by µt = d<br />

dt (log λt) ◦ ψ −1<br />

t , cf. the proof of the Gray<br />

stability theorem (Theorem 2.20).<br />

Write Xt = HtRt+Yt with Rt the Reeb vector field of αt and Yt ∈ ξt = kerαt.<br />

Then condition (2.8) translates into<br />

˙αt + dHt + iYtdαt = µtαt. (2.9)<br />

For given Ht one determines µt from this equation by inserting the Reeb vector<br />

field Rt; the equation then admits a unique solution Yt ∈ kerαt because of the<br />

non-degeneracy of dαt|ξt .<br />

Our aim now is to ensure that Ht ≡ 0 on S0 and Yt ≡ 0 along S0. The latter<br />

we achieve by imposing the condition<br />

˙αt + dHt = 0 along S0<br />

(2.10)<br />

(which entails with (2.9) that µt|S0 ≡ 0). The conditions Ht ≡ 0 on S0 and (2.10)<br />

can be simultaneously satisfied thanks to ˙αt|TS0 ≡ 0.<br />

We can therefore find a smooth family of smooth functions Ht satisfying these<br />

conditions, and then define Yt by (2.9). The flow of the vector field Xt = HtRt+Yt<br />

then defines an isotopy ψt that fixes S0 pointwise (and thus is defined for all<br />

t ∈ [0, 1] in a neighbourhood of S0). Then ψ = φ ◦ ψ1 will be the diffeomorphism<br />

we wanted to construct.<br />

2.4.5 Applications<br />

Perhaps the most important consequence of the neighbourhood theorems proved<br />

above is that they allow us to perform differential topological constructions such<br />

30


as surgery or similar cutting and pasting operations in the presence of a contact<br />

structure, that is, these constructions can be carried out on a contact manifold<br />

in such a way that the resulting manifold again carries a contact structure.<br />

One such construction that I shall explain in detail in Section 3 is the surgery<br />

of contact 3–manifolds along transverse knots, which enables us to construct a<br />

contact structure on every closed, orientable 3–manifold.<br />

The concept of characteristic foliation on surfaces in contact 3–manifolds<br />

has proved seminal for the classification of contact structures on 3–manifolds,<br />

although it has recently been superseded by the notion of dividing curves. It is<br />

clear that Theorem 2.39 can be used to cut and paste contact manifolds along<br />

hypersurfaces with the same characteristic foliation. What actually makes this<br />

useful in dimension 3 is that there are ways to manipulate the characteristic<br />

foliation of a surface by isotoping that surface inside the contact 3–manifold.<br />

The most important result in that direction is the Elimination Lemma proved<br />

by Giroux [52]; an improved version is due to D. Fuchs, see [26]. This lemma<br />

says that un<strong>der</strong> suitable assumptions it is possible to cancel singular points of the<br />

characteristic foliation in pairs by a C 0 –small isotopy of the surface (specifically:<br />

an elliptic and a hyperbolic point of the same sign – the sign being determined<br />

by the matching or non-matching of the orientation of the surface S and the<br />

contact structure ξ at the singular point of Sξ). For instance, Eliashberg [24] has<br />

shown that if a contact 3–manifold (M, ξ) contains an embedded disc D ′ such<br />

that D ′ ξ<br />

has a limit cycle, then one can actually find a so-called overtwisted disc:<br />

an embedded disc D with boundary ∂D tangent to ξ (but D transverse to ξ<br />

along ∂D, i.e. no singular points of Dξ on ∂D) and with Dξ having exactly one<br />

singular point (of elliptic type); see Section 3.6.<br />

Moreover, in the generic situation it is possible, given surfaces S ⊂ (M, ξ)<br />

and S ′ ⊂ (M ′ , ξ ′ ) with Sξ homeomorphic to S ′ ξ ′, to perturb one of the surfaces so<br />

as to get diffeomorphic characteristic foliations.<br />

Chapter 8 of [1] contains a section on surfaces in contact 3–manifolds, and<br />

in particular a proof of the Elimination Lemma. Further introductory reading<br />

on the matter can be found in the lectures of J. Etnyre [35]; of the sources cited<br />

above I recommend [26] as a starting point.<br />

In [52] Giroux initiated the study of convex surfaces in contact 3–manifolds.<br />

These are surfaces S with an infinitesimal automorphism X of the contact struc-<br />

ture ξ with X transverse to S. For such surfaces, it turns out, much less infor-<br />

31


mation than the characteristic foliation Sξ is needed to determine ξ in a neigh-<br />

bourhood of S, viz., only the so-called dividing set of Sξ. This notion lies at the<br />

centre of most of the recent advances in the classification of contact structures<br />

on 3–manifolds [55], [71], [72]. A brief introduction to convex surface theory can<br />

be found in [35].<br />

2.5 Isotopy extension theorems<br />

In this section we show that the isotopy extension theorem of differential topology<br />

– an isotopy of a closed submanifold extends to an isotopy of the ambient manifold<br />

– remains valid for the various distinguished submanifolds of contact manifolds.<br />

The neighbourhood theorems proved above provide the key to the corresponding<br />

isotopy extension theorems. For simplicity, I assume throughout that the ambient<br />

contact manifold M is closed; all isotopy extension theorems remain valid if M has<br />

non-empty boundary ∂M, provided the isotopy stays away from the boundary.<br />

In that case, the isotopy of M found by extension keeps a neighbourhood of<br />

∂M fixed. A further convention throughout is that our ambient isotopies ψt are<br />

un<strong>der</strong>stood to start at ψ0 = idM.<br />

2.5.1 Isotropic submanifolds<br />

An embedding j : L → (M, ξ = kerα) is called isotropic if j(L) is an isotropic<br />

submanifold of (M, ξ), i.e. everywhere tangent to the contact structure ξ. Equiv-<br />

alently, one needs to require j ∗ α ≡ 0.<br />

Theorem 2.41. Let jt: L → (M, ξ = kerα), t ∈ [0, 1], be an isotopy of isotropic<br />

embeddings of a closed manifold L in a contact manifold (M, ξ). Then there is a<br />

compactly supported contact isotopy ψt: M → M with ψt(j0(L)) = jt(L).<br />

Proof. Define a time-dependent vector field Xt along jt(L) by<br />

Xt ◦ jt = d<br />

dt jt.<br />

To simplify notation later on, we assume that L is a submanifold of M and j0 the<br />

inclusion L ⊂ M. Our aim is to find a (smooth) family of compactly supported,<br />

smooth functions � Ht: M → R whose Hamiltonian vector field � Xt equals Xt along<br />

jt(L). Recall that � Xt is defined in terms of � Ht by<br />

α( � Xt) = � Ht, i�Xt dα = d � Ht(Rα)α − d � Ht,<br />

32


where, as usual, Rα denotes the Reeb vector field of α.<br />

Hence, we need<br />

α(Xt) = � Ht, iXtdα = d � Ht(Rα)α − d � Ht along jt(L). (2.11)<br />

Write Xt = HtRα + Yt with Ht: jt(L) → R and Yt ∈ kerα. To satisfy (2.11) we<br />

need<br />

This implies<br />

�Ht = Ht along jt(L). (2.12)<br />

d � Ht(v) = dHt(v) for v ∈ T(jt(L)).<br />

Since jt is an isotopy of isotropic embeddings, we have T(jt(L)) ⊂ ker α. So a<br />

prerequisite for (2.11) is that<br />

We have<br />

dα(Xt, v) = −dHt(v) for v ∈ T(jt(L)). (2.13)<br />

dα(Xt, v) + dHt(v) = dα(Xt, v) + d(α(Xt))(v)<br />

so equation (2.13) is equivalent to<br />

= iv(iXtdα + d(iXtα))<br />

= iv(LXtα),<br />

iv(LXtα) = 0 for v ∈ T(jt(L)).<br />

But this is indeed tautologically satisfied: The fact that jt is an isotopy of isotropic<br />

embeddings can be written as j ∗ t α ≡ 0; this in turn implies 0 = d<br />

dt (j∗ t α) =<br />

j ∗ t (LXtα), as desired.<br />

This means that we can define � Ht by prescribing the value of � Ht along jt(L)<br />

(with (2.12)) and the differential of � Ht along jt(L) (with (2.11)), where we are<br />

free to impose d � Ht(Rα) = 0, for instance. The calculation we just performed<br />

shows that these two requirements are consistent with each other. Any function<br />

satisfying these requirements along jt(L) can be smoothed out to zero outside a<br />

tubular neighbourhood of jt(L), and the Hamiltonian flow of this � Ht will be the<br />

desired contact isotopy extending jt.<br />

One small technical point is to ensure that the resulting family of functions<br />

�Ht will be smooth in t. To achieve this, we proceed as follows. Set ˆ M = M ×[0, 1]<br />

and<br />

ˆL = �<br />

q∈L,t∈[0,1]<br />

33<br />

(jt(q), t),


so that ˆ L is a submanifold of ˆ M. Let g be an auxiliary Riemannian metric on M<br />

with respect to which Rα is orthogonal to kerα. Identify the normal bundle N ˆ L<br />

of ˆ L in ˆ M with a sub-bundle of T ˆ M by requiring its fibre at a point (p, t) ∈ ˆ L<br />

to be the g–orthogonal subspace of Tp(jt(L)) in TpM. Let τ : N ˆ L → ˆ M be a<br />

tubular map.<br />

Now define a smooth function ˆ H : N ˆ L → R as follows, where (p, t) always<br />

denotes a point of ˆ L ⊂ N ˆ L.<br />

• ˆ H(p, t) = α(Xt),<br />

• d ˆ H (p,t)(Rα) = 0,<br />

• d ˆ H (p,t)(v) = −dα(Xt, v) for v ∈ kerαp ⊂ TpM ⊂ T (p,t) ˆ M,<br />

• ˆ H is linear on the fibres of N ˆ L → ˆ L.<br />

Let χ: ˆ M → [0, 1] be a smooth function with χ ≡ 0 outside a small neighbour-<br />

hood N0 ⊂ τ(N ˆ L) of ˆ L and χ ≡ 1 in a smaller neighbourhood N1 ⊂ N0 of ˆ L.<br />

For (p, t) ∈ ˆ M, set<br />

�Ht(p) =<br />

�<br />

χ(p, t) ˆ H(τ −1 (p, t)) for (p, t) ∈ τ(N ˆ L)<br />

0 for (p, t) �∈ τ(N ˆ L).<br />

This is smooth in p and t, and the Hamiltonian flow ψt of � Ht (defined globally<br />

since � Ht is compactly supported) is the desired contact isotopy.<br />

2.5.2 <strong>Contact</strong> submanifolds<br />

An embedding j : (M ′ , ξ ′ ) → (M, ξ) is called a contact embedding if<br />

is a contact submanifold of (M, ξ), i.e.<br />

(j(M ′ ), Tj(ξ ′ ))<br />

T(j(M)) ∩ ξ| j(M) = Tj(ξ ′ ).<br />

If ξ = kerα, this can be reformulated as kerj ∗ α = ξ ′ .<br />

Theorem 2.42. Let jt: (M ′ , ξ ′ ) → (M, ξ), t ∈ [0, 1], be an isotopy of con-<br />

tact embeddings of the closed contact manifold (M ′ , ξ ′ ) in the contact manifold<br />

(M, ξ). Then there is a compactly supported contact isotopy ψt: M → M with<br />

ψt(j0(M ′ )) = jt(M ′ ).<br />

34


Proof. In the proof of this theorem we follow a slightly different strategy from<br />

the one in the isotropic case. Instead of directly finding an extension of the<br />

Hamiltonian Ht: jt(M ′ ) → R, we first use the neighbourhood theorem for con-<br />

tact submanifolds to extend jt to an isotopy of contact embeddings of tubular<br />

neighbourhoods.<br />

Again we assume that M ′ is a submanifold of M and j0 the inclusion M ′ ⊂ M.<br />

As earlier, NM ′ denotes the normal bundle of M ′ in M. We also identify M ′<br />

with the zero section of NM ′ , and we use the canonical identification<br />

T(NM ′ )|M ′ = TM ′ ⊕ NM ′ .<br />

By the usual isotopy extension theorem from differential topology we find an<br />

isotopy<br />

φt: NM ′ −→ M<br />

with φt|M ′ = jt.<br />

Choose contact forms α, α ′ defining ξ and ξ ′ , respectively. Define αt = φ∗ tα.<br />

Then TM ′ ∩ kerαt = ξ ′ . Let R ′ denote the Reeb vector field of α ′ . Analogous<br />

to the proof of Theorem 2.32, we first find a smooth family of smooth functions<br />

gt: M ′ → R + such that gtαt|TM ′ = α′ , and then a family ft: NM ′ → R + with<br />

ft|M ′ ≡ 1 and<br />

dft = iR ′d(gtαt) on T(NM ′ )|M ′.<br />

Then βt = ftgtαt is a family of contact forms on NM ′ representing the contact<br />

structure ker(φ ∗ tα) and with the properties<br />

βt|TM ′ = α′ ,<br />

dβt|TM ′ = dα′ ,<br />

ker(dβt) = 〈R ′ 〉 along M ′ .<br />

The family (NM ′ , dβt) of symplectic vector bundles may be thought of as a<br />

symplectic vector bundle over M ′ × [0, 1], which is necessarily isomorphic to a<br />

bundle pulled back from M ′ × {0} (cf. [74, Cor. 3.4.4]). In other words, there is<br />

a smooth family of symplectic bundle isomorphisms<br />

Then<br />

Φt: (NM ′ , dβ0) −→ (NM ′ , dβt).<br />

idTM ′ ⊕ Φt: T(NM ′ )|M ′ −→ T(NM′ )|M ′<br />

35


is a bundle map that pulls back βt to β0 and dβt to dβ0.<br />

By the now familiar stability argument we find a smooth family of embeddings<br />

ϕt: N(M ′ ) −→ NM ′<br />

for some neighbourhood N(M ′ ) of the zero section M ′ in NM ′ with ϕ0 =<br />

inclusion, ϕt|M ′ = idM ′ and ϕ∗ tβt = λtβ0, where λt: N(M ′ ) → R + . This means<br />

that<br />

φt ◦ ϕt: N(M ′ ) −→ M<br />

is a smooth family of contact embeddings of (N(M ′ ), ker β0) in (M, ξ).<br />

Define a time-dependent vector field Xt along φt ◦ ϕt(N(M ′ )) by<br />

Xt ◦ φt ◦ ϕt = d<br />

dt (φt ◦ ϕt).<br />

This Xt is clearly an infinitesimal automorphism of ξ: By differentiating the<br />

equation ϕ ∗ tφ ∗ tα = µtφ ∗ 0 α (where µt: N(M ′ ) → R + ) with respect to t we get<br />

ϕ ∗ tφ ∗ t(LXtα) = ˙µtφ ∗ 0α = ˙µt<br />

ϕ ∗ tφ ∗ tα,<br />

so LXtα is a multiple of α (since φt ◦ ϕt is a diffeomorphism onto its image).<br />

By the theory of contact Hamiltonians, Xt is the Hamiltonian vector field of<br />

a Hamiltonian function ˆ Ht defined on φt ◦ ϕt(N(M ′ )). Cut off this function with<br />

a bump function so as to obtain Ht: M → R with Ht ≡ ˆ Ht near φt ◦ ϕt(M ′ )<br />

and Ht ≡ 0 outside a slightly larger neighbourhood of φt ◦ ϕt(M ′ ). Then the<br />

Hamiltonian flow ψt of Ht satisfies our requirements.<br />

2.5.3 Surfaces in 3–manifolds<br />

Theorem 2.43. Let jt: S → (M, ξ = kerα), t ∈ [0, 1], be an isotopy of embed-<br />

dings of a closed surface S in a 3–dimensional contact manifold (M, ξ). If all jt<br />

induce the same characteristic foliation on S, then there is a compactly supported<br />

isotopy ψt: M → M with ψt(j0(S)) = jt(S).<br />

Proof. Extend jt to a smooth family of embeddings φt: S ×R → M, and identify<br />

S with S × {0}. The assumptions say that all φ ∗ tα induce the same characteristic<br />

foliation on S. By the proof of Theorem 2.40 and in analogy with the proof of<br />

Theorem 2.42 we find a smooth family of embeddings<br />

µt<br />

ϕt: S × (−ε, ε) −→ S × R<br />

36


for some ε > 0 with ϕ0 = inclusion, ϕt| S×{0} = idS and ϕ∗ tφ∗ tα = λtφ∗ 0α, where<br />

λt: S × (−ε, ε) → R + . In other words, φt ◦ ϕt is a smooth family of contact<br />

embeddings of (S × (−ε, ε), ker φ∗ 0α) in (M, ξ).<br />

The proof now concludes exactly as the proof of Theorem 2.42.<br />

2.6 Approximation theorems<br />

A further manifestation of the (local) flexibility of contact structures is the fact<br />

that a given submanifold can, un<strong>der</strong> fairly weak (and usually obvious) topological<br />

conditions, be approximated (typically C 0 –closely) by a contact submanifold or<br />

an isotropic submanifold, respectively. The most general results in this direction<br />

are best phrased in M. Gromov’s language of h-principles. For a recent text on<br />

h-principles that puts particular emphasis on symplectic and contact geometry<br />

see [30]; a brief and perhaps more gentle introduction to h-principles can be found<br />

in [47].<br />

In the present section I restrict attention to the 3–dimensional situation, where<br />

the relevant approximation theorems can be proved by elementary geometric ad<br />

hoc techniques.<br />

Theorem 2.44. Let γ : S 1 → (M, ξ) be a knot, i.e. an embedding of S 1 , in<br />

a contact 3–manifold. Then γ can be C 0 –approximated by a Legendrian knot<br />

isotopic to γ. Alternatively, it can be C 0 –approximated by a positively as well as<br />

a negatively transverse knot.<br />

In or<strong>der</strong> to prove this theorem, we first consi<strong>der</strong> embeddings γ : (a, b) →<br />

(R 3 , ξ) of an open interval in R 3 with its standard contact structure ξ = kerα,<br />

where α = dz + x dy. Write γ(t) = (x(t), y(t), z(t)). Then<br />

α(˙γ) = ˙z + x ˙y,<br />

so the condition for a Legendrian curve reads ˙z + x ˙y ≡ 0; for a positively (resp.<br />

negatively) transverse curve, ˙z + x ˙y > 0 (resp. < 0).<br />

There are two ways to visualise such curves. The first is via its front pro-<br />

jection<br />

γF(t) = (y(t), z(t)),<br />

the second via its Lagrangian projection<br />

γL(t) = (x(t), y(t)).<br />

37


2.6.1 Legendrian knots<br />

If γ(t) = (x(t), y(t), z(t)) is a Legendrian curve in R 3 , then ˙y = 0 implies ˙z = 0,<br />

so there the front projection has a singular point. Indeed, the curve t ↦→ (t, 0, 0)<br />

is an example of a Legendrian curve whose front projection is a single point. We<br />

call a Legendrian curve generic if ˙y = 0 only holds at isolated points (which we<br />

call cusp points), and there ¨y �= 0.<br />

Lemma 2.45. Let γ : (a, b) → (R 3 , ξ) be a Legendrian immersion. Then its front<br />

projection γF(t) = (y(t), z(t)) does not have any vertical tangencies. Away from<br />

the cusp points, γ is recovered from its front projection via<br />

x(t) = − ˙z(t)<br />

= −dz<br />

˙y(t) dy ,<br />

i.e. x(t) is the negative slope of the front projection. The curve γ is embedded if<br />

and only if γF has only transverse self-intersections.<br />

By a C ∞ –small perturbation of γ we can obtain a generic Legendrian curve ˜γ<br />

isotopic to γ; by a C 2 –small perturbation we may achieve that the front projection<br />

has only semi-cubical cusp singularities, i.e. around a cusp point at t = 0 the curve<br />

˜γ looks like<br />

with λ �= 0, see Figure 3.<br />

˜γ(t) = (t + a, λt 2 + b, −λ(2t 3 /3 + at 2 ) + c)<br />

Any regular curve in the (y, z)–plane with semi-cubical cusps and no vertical<br />

tangencies can be lifted to a unique Legendrian curve in R 3 .<br />

Figure 3: The cusp of a front projection.<br />

Proof. The Legendrian condition is ˙z + x ˙y = 0. Hence ˙y = 0 forces ˙z = 0, so γF<br />

cannot have any vertical tangencies.<br />

Away from the cusp points, the Legendrian condition tells us how to recover<br />

x as the negative slope of the front projection. (By continuity, the equation<br />

38


x = dz<br />

dy<br />

also makes sense at generic cusp points.) In particular, a self-intersecting<br />

front projection lifts to a non-intersecting curve if and only if the slopes at the<br />

intersection point are different, i.e. if and only if the intersection is transverse.<br />

That γ can be approximated in the C ∞ –topology by a generic immersion<br />

˜γ follows from the usual transversality theorem (in its most simple form, viz.,<br />

applied to the function y(t); the function x(t) may be left unchanged, and the<br />

new z(t) is then found by integrating the new −x ˙y).<br />

At a cusp point of ˜γ we have ˙y = ˙z = 0. Since ˜γ is an immersion, this forces<br />

˙x �= 0, so ˜γ can be parametrised around a cusp point by the x–coordinate, i.e. we<br />

may choose the curve parameter t such that the cusp lies at t = 0 and x(t) = t+a.<br />

Since ¨y(0) �= 0 by the genericity condition, we can write y(t) = t 2 g(t) + y(0)<br />

with a smooth function g(t) satisfying g(0) �= 0 (This is proved like the ‘Morse<br />

lemma’ in Appendix 2 of [77]). A C 0 –approximation of g(t) by a function h(t)<br />

with h(t) ≡ g(0) for t near zero and h(t) ≡ g(t) for |t| greater than some small<br />

ε > 0 yields a C 2 –approximation of y(t) with the desired form around the cusp<br />

point.<br />

Example 2.46. Figure 4 shows the front projection of a Legendrian trefoil knot.<br />

Figure 4: Front projection of a Legendrian trefoil knot.<br />

Proof of Theorem 2.44 - Legendrian case. First of all, we consi<strong>der</strong> a curve γ in<br />

standard R 3 . In or<strong>der</strong> to find a C 0 –close approximation of γ, we simply need<br />

to choose a C 0 –close approximation of its front projection γF by a regular curve<br />

without vertical tangencies and with isolated cusps (we call such a curve a front)<br />

39


in such a way, that the slope of the front at time t is close to −x(t) (see Figure 5).<br />

Then the Legendrian lift of this front is the desired C 0 –approximation of γ.<br />

z<br />

Figure 5: Legendrian C 0 –approximation via front projection.<br />

If γ is defined on an interval (a, b) and is already Legendrian near its endpoints,<br />

then the approximation of γF may be assumed to coincide with γF near the<br />

endpoints, so that the Legendrian lift coincides with γ near the endpoints.<br />

Hence, given a knot in an arbitrary contact 3–manifold, we can cut it (by the<br />

Lebesgue lemma) into little pieces that lie in Darboux charts. There we can use<br />

the preceding recipe to find a Legendrian approximation. Since, as just observed,<br />

one can find such approximations on intervals with given boundary condition,<br />

this procedure yields a Legendrian approximation of the full knot.<br />

Locally (i.e. in R 3 ) the described procedure does not introduce any self-<br />

intersections in the approximating curve, provided we approximate γF by a front<br />

with only transverse self-intersections. Since the original knot was embedded, the<br />

same will then be true for its Legendrian C 0 –approximation.<br />

The same result may be <strong>der</strong>ived using the Lagrangian projection:<br />

Lemma 2.47. Let γ : (a, b) → (R 3 , ξ) be a Legendrian immersion. Then its<br />

Lagrangian projection γL(t) = (x(t), y(t)) is also an immersed curve. The curve<br />

γ is recovered from γL via<br />

� t1<br />

z(t1) = z(t0) −<br />

40<br />

t0<br />

x dy.<br />

y


A Legendrian immersion γ : S 1 → (R 3 , ξ) has a Lagrangian projection that en-<br />

closes zero area. Moreover, γ is embedded if and only if every loop in γL (except,<br />

in the closed case, the full loop γL) encloses a non-zero oriented area.<br />

Any immersed curve in the (x, y)–plane is the Lagrangian projection of a<br />

Legendrian curve in R 3 , unique up to translation in the z–direction.<br />

Proof. The Legendrian condition ˙z + x ˙y implies that if ˙y = 0 then ˙z = 0, and<br />

hence, since γ is an immersion, ˙x �= 0. So γL is an immersion.<br />

The formula for z follows by integrating the Legendrian condition. For a<br />

closed curve γL in the (x, y)–plane, the integral �<br />

x dy computes the oriented<br />

area enclosed by γL. From that all the other statements follow.<br />

Example 2.48. Figure 6 shows the Lagrangian projection of a Legendrian un-<br />

knot.<br />

A −2A<br />

Figure 6: Lagrangian projection of a Legendrian unknot.<br />

Alternative proof of Theorem 2.44 – Legendrian case. Again we consi<strong>der</strong> a curve<br />

γ in standard R 3 defined on an interval. The generalisation to arbitrary contact<br />

manifolds and closed curves is achieved as in the proof using front projections.<br />

In or<strong>der</strong> to find a C 0 –approximation of γ by a Legendrian curve, one only has<br />

to approximate its Lagrangian projection γL by an immersed curve whose ‘area<br />

integral’<br />

z(t0) −<br />

lies as close to the original z(t) as one wishes. This can be achieved by using small<br />

loops oriented positively or negatively (see Figure 7). If γL has self-intersections,<br />

this approximating curve can be chosen in such a way that along loops properly<br />

contained in that curve the area integral is non-zero, so that again we do not<br />

� t<br />

t0<br />

x dy<br />

introduce any self-intersections in the Legendrian approximation of γ.<br />

41<br />

γL<br />

A


y<br />

Figure 7: Legendrian C 0 –approximation via Lagrangian projection.<br />

2.6.2 Transverse knots<br />

The quickest proof of the transverse case of Theorem 2.44 is via the Legendrian<br />

case. However, it is perfectly feasible to give a direct proof along the lines of the<br />

preceding discussion, i.e. using the front or the Lagrangian projection. Since this<br />

picture is useful elsewhere, I include a brief discussion, restricting attention to<br />

the front projection.<br />

Thus, let γ(t) = (x(t), y(t), z(t)) be a curve in R 3 . The condition for γ to be<br />

positively transverse to the standard contact structure ξ = ker(dz + x dy) is that<br />

˙z + x ˙y > 0. Hence,<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

if ˙y = 0, then ˙z > 0,<br />

if ˙y > 0, then x > − ˙z/ ˙y,<br />

if ˙y < 0, then x < − ˙z/ ˙y.<br />

The first statement says that there are no vertical tangencies oriented down-<br />

wards in the front projection. The second statement says in particular that for<br />

˙y > 0 and ˙z < 0 we have x > 0; the third, that for ˙y < 0 and ˙z < 0 we have<br />

x < 0. This implies that the situations shown in Figure 8 are not possible in the<br />

front projection of a positively transverse curve. I leave it to the rea<strong>der</strong> to check<br />

that all other oriented crossings are possible in the front projection of a positively<br />

transverse curve, and that any curve in the (y, z)–plane without the forbidden<br />

crossing or downward vertical tangencies admits a lift to a positively transverse<br />

curve.<br />

42<br />

x


Figure 8: Impossible front projections of positively transverse curve.<br />

Example 2.49. Figure 9 shows the front projection of a positively transverse<br />

trefoil knot.<br />

Figure 9: Front projection of a positively transverse trefoil knot.<br />

Proof of Theorem 2.44 – transverse case. By the Legendrian case of this theo-<br />

rem, the given knot γ can be C 0 –approximated by a Legendrian knot γ1. By<br />

Example 2.29, a neighbourhood of γ1 in (M, ξ) looks like a solid torus S 1 × D 2<br />

with contact structure cosθ dx − sinθ dy = 0, where γ1 ≡ S 1 × {0}. Then the<br />

curve<br />

γ2(t) = (θ = t, x = δ sint, y = δ cos t), t ∈ [0, 2π],<br />

is a positively (resp. negatively) transverse knot if δ > 0 (resp. < 0). By choosing<br />

|δ| small we obtain as good a C 0 –approximation of γ1 and hence of γ as we<br />

wish.<br />

3 <strong>Contact</strong> structures on 3–manifolds<br />

Here is the main theorem proved in this section:<br />

43


Theorem 3.1 (Lutz-Martinet). Every closed, orientable 3–manifold admits a<br />

contact structure in each homotopy class of tangent 2–plane fields.<br />

In Section 3.2 I present what is essentially J. Martinet’s [90] proof of the<br />

existence of a contact structure on every 3–manifold. This construction is based<br />

on a surgery description of 3–manifolds due to R. Lickorish and A. Wallace. For<br />

the key step, showing how to extend over a solid torus certain contact forms<br />

defined near the boundary of that torus (which then makes it possible to perform<br />

Dehn surgeries), we use an approach due to W. Thurston and H. Winkelnkemper;<br />

this allows to simplify Martinet’s proof slightly.<br />

In Section 3.3 we show that every orientable 3–manifold is parallelisable and<br />

then build on this to classify (co-)oriented tangent 2–plane fields up to homotopy.<br />

In Section 3.4 we study the so-called Lutz twist, a topologically trivial Dehn<br />

surgery on a contact manifold (M, ξ) which yields a contact structure ξ ′ on M<br />

that is not homotopic (as 2–plane field) to ξ. We then complete the proof of the<br />

main theorem stated above. These results are contained in R. Lutz’s thesis [84]<br />

(which, I have to admit, I’ve never seen). Of Lutz’s published work, [83] only deals<br />

with the 3–sphere (and is only an announcement); [85] also deals with a more<br />

restricted problem. I learned the key steps of the construction from an exposition<br />

given in V. Ginzburg’s thesis [50]. I have added proofs of many topological details<br />

that do not seem to have appeared in a readily accessible source before.<br />

In Section 3.5 I indicate two further proofs for the existence of contact struc-<br />

tures on every 3–manifold (and provide references to others). The one by Thur-<br />

ston and Winkelnkemper uses a description of 3–manifolds as open books due to<br />

J. Alexan<strong>der</strong>; the crucial idea in their proof is the one we also use to simplify<br />

Martinet’s argument. Indeed, my discussion of the Lutz twist in the present<br />

section owes more to the paper by Thurston-Winkelnkemper than to any other<br />

reference. The second proof, by J. Gonzalo, is based on a branched cover descrip-<br />

tion of 3–manifolds found by H. Hilden, J. Montesinos and T. Thickstun. This<br />

branched cover description also yields a very simple geometric proof that every<br />

orientable 3–manifold is parallelisable.<br />

In Section 3.6 we discuss the fundamental dichotomy between tight and over-<br />

twisted contact structures, introduced by Eliashberg, as well as the relation of<br />

these types of contact structures with the concept of symplectic fillability. The<br />

chapter concludes in Section 3.7 with a survey of classification results for contact<br />

structures on 3–manifolds.<br />

44


But first we discuss, in Section 3.1, an invariant of transverse knots in R 3<br />

with its standard contact structure. This invariant will be an ingredient in the<br />

proof of the Lutz-Martinet theorem, but is also of independent interest.<br />

I do not feel embarrassed to use quite a bit of machinery from algebraic and<br />

differential topology in this chapter. However, I believe that nothing that cannot<br />

be found in such standard texts as [14], [77] and [95] is used without proof or an<br />

explicit reference.<br />

Throughout this third section, M denotes a closed, orientable 3-manifold.<br />

3.1 An invariant of transverse knots<br />

Although the invariant in question can be defined for transverse knots in arbitrary<br />

contact manifolds (provided the knot is homologically trivial), for the sake of<br />

clarity I restrict attention to transverse knots in R 3 with its standard contact<br />

structure ξ0 = ker(dz + x dy). This will be sufficient for the proof of the Lutz-<br />

Martinet theorem. In Section 3.7 I say a few words about the general situation<br />

and related invariants for Legendrian knots.<br />

Thus, let γ be a transverse knot in (R 3 , ξ0). Push γ a little in the direction<br />

of ∂x – notice that this is a nowhere zero vector field contained in ξ0, and in<br />

particular transverse to γ – to obtain a knot γ ′ . An orientation of γ induces an<br />

orientation of γ ′ . The orientation of R 3 is given by dx ∧ dy ∧ dz.<br />

Definition 3.2. The self-linking number l(γ) of the transverse knot γ is the<br />

linking number of γ and γ ′ .<br />

Notice that this definition is independent of the choice of orientation of γ (but<br />

it changes sign if the orientation of R 3 is reversed). Furthermore, in place of ∂x<br />

we could have chosen any nowhere zero vector field X in ξ0 to define l(γ): The<br />

difference between the the self-linking number defined via ∂x and that defined<br />

via X is the degree of the map γ → S 1 given by associating to a point on γ<br />

the angle between ∂x and X. This degree is computed with the induced map<br />

Z ∼ = H1(γ) → H1(S 1 ) ∼ = Z. But the map γ → S 1 factors through R 3 , so the<br />

induced homomorphism on homology is the zero homomorphism.<br />

Observe that l(γ) is an invariant un<strong>der</strong> isotopies of γ within the class of<br />

transverse knots.<br />

We now want to compute l(γ) from the front projection of γ. Recall that the<br />

writhe of an oriented knot diagram is the signed number of self-crossings of the<br />

45


diagram, where the sign of the crossing is given in Figure 10.<br />

−1<br />

+1<br />

Figure 10: Signs of crossings in a knot diagram.<br />

Lemma 3.3. The self-linking number l(γ) of a transverse knot is equal to the<br />

writhe w(γ) of its front projection.<br />

Proof. Let γ ′ be the push-off of γ as described. Observe that each crossing of the<br />

front projection of γ contributes a crossing of γ ′ un<strong>der</strong>neath γ of the corresponding<br />

sign. Since the linking number of γ and γ ′ is equal to the signed number of times<br />

that γ ′ crosses un<strong>der</strong>neath γ (cf. [98, p. 37]), we find that this linking number is<br />

equal to the signed number of self-crossings of γ, that is, l(γ) = w(γ).<br />

Proposition 3.4. Every self-linking number is realised by a transverse link in<br />

standard R 3 .<br />

Proof. Figure 11 shows front projections of positively transverse knots (cf. Sec-<br />

tion 2.6.2) with self-linking number ±3. From that the construction principle for<br />

realising any odd integer should be clear. With a two component link any even<br />

integer can be realised.<br />

Remark 3.5. It is no accident that I do not give an example of a transverse<br />

knot with even self-linking number. By a theorem of Eliashberg [26, Prop. 2.3.1]<br />

that relates l(γ) to the Euler characteristic of a Seifert surface S for γ and the<br />

signed number of singular points of the characteristic foliation Sξ, the self-linking<br />

number l(γ) of a knot can only take odd values.<br />

3.2 Martinet’s construction<br />

According to Lickorish [81] and Wallace [103] M can be obtained from S 3 by<br />

Dehn surgery along a link of 1–spheres. This means that we have to remove<br />

46


−3<br />

+3<br />

Figure 11: Transverse knots with self-linking number ±3.<br />

a disjoint set of embedded solid tori S 1 × D 2 from S 3 and glue back solid tori<br />

with suitable identification by a diffeomorphism along the boundaries S 1 × S 1 .<br />

The effect of such a surgery (up to diffeomorphism of the resulting manifold) is<br />

completely determined by the induced map in homology<br />

H1(S 1 × ∂D 2 ) −→ H1(S 1 × ∂D 2 )<br />

Z ⊕ Z −→ Z ⊕ Z,<br />

� �<br />

which is given by a unimodular matrix<br />

n q<br />

m p<br />

∈ GL(2, Z). Hence, denoting<br />

coordinates in S 1 × S 1 by (θ, ϕ), we may always assume the identification maps<br />

to be of the form �<br />

θ<br />

ϕ<br />

�<br />

↦−→<br />

�<br />

n q<br />

m p<br />

The curves µ and λ on ∂(S 1 × D 2 ) given respectively by θ = 0 and ϕ = 0 are<br />

called meridian and longitude. We keep the same notation µ, λ for the homology<br />

classes these curves represent. It turns out that the diffeomorphism type of the<br />

surgered manifold is completely determined by the class pµ + qλ, which is the<br />

class of the curve that becomes homotopically trivial in the surgered manifold<br />

(cf. [98, p. 28]). In fact, the Dehn surgery is completely determined by the surgery<br />

coefficient p/q, since the diffeomorphism of ∂(S 1 ×D 2 ) given by (λ, µ) ↦→ (λ, −µ)<br />

��<br />

extends to a diffeomorphism of the solid torus that we glue back.<br />

Similarly, the diffeomorphism given by (λ, µ) ↦→ (λ + kµ, µ) extends. By such<br />

a change of longitude in S 1 × D 2 ⊂ M, which amounts to choosing a different<br />

47<br />

θ<br />

ϕ<br />

�<br />

.


trivialisation of the normal bundle (i.e. framing) of S1 �<br />

�<br />

×{0} ⊂ M, the gluing map<br />

is changed to<br />

n q<br />

m − kn p − kq<br />

. By a change of longitude in the solid torus<br />

� �<br />

that we glue back, the gluing map is changed to<br />

n + kq q<br />

m + kp p<br />

. Thus, a Dehn<br />

surgery is a so-called handle surgery (or ‘honest surgery’ or simply ‘surgery’) if<br />

and only if the surgery coefficient � is �an<br />

integer, � that � is, q = ±1. For in exactly<br />

this case we may assume<br />

n q<br />

m p<br />

=<br />

0 1<br />

1 0<br />

, and the surgery is given by<br />

cutting out S 1 × D 2 and gluing back S 1 × D 2 with the identity map<br />

∂(D 2 × S 1 ) −→ ∂(S 1 × D 2 ).<br />

The theorem of Lickorish and Wallace remains true if we only allow handle<br />

surgeries. However, this assumption does not entail any great simplification of<br />

the existence proof for contact structures, so we shall work with general Dehn<br />

surgeries.<br />

Our aim in this section is to use this topological description of 3–manifolds<br />

for a proof of the following theorem, first proved by Martinet [90]. The proof<br />

presented here is in spirit the one given by Martinet, but, as indicated in the<br />

introduction to this third section, amalgamated with ideas of Thurston and<br />

Winkelnkemper [101], whose proof of the same theorem we shall discuss later.<br />

Theorem 3.6 (Martinet). Every closed, orientable 3–manifold M admits a con-<br />

tact structure.<br />

In view of the theorem of Lickorish and Wallace and the fact that S 3 admits<br />

a contact structure, Martinet’s theorem is a direct consequence of the following<br />

result.<br />

Theorem 3.7. Let ξ0 be a contact structure on a 3–manifold M0. Let M be the<br />

manifold obtained from M0 by a Dehn surgery along a knot K. Then M admits a<br />

contact structure ξ which coincides with ξ0 outside the neighbourhood of K where<br />

we perform surgery.<br />

Proof. By Theorem 2.44 we may assume that K is positively transverse to ξ0.<br />

Then, by the contact neighbourhood theorem (Example 2.33), we can find a<br />

tubular neighbourhood of K diffeomorphic to S 1 × D 2 (δ0), where K is identified<br />

with S 1 × {0} and D 2 (δ0) denotes a disc of radius δ0, such that the contact<br />

48


structure ξ0 is given as the kernel of dθ+r 2 dϕ, with θ denoting the S 1 –coordinate<br />

and (r, ϕ) polar coordinates on D 2 (δ0).<br />

� Now perform � a Dehn surgery along K defined by the unimodular matrix<br />

n q<br />

m p<br />

. This corresponds to cutting out S 1 ×D 2 (δ1) ⊂ S 1 ×D 2 (δ0) for some<br />

δ1 < δ0 and gluing it back in the way described above.<br />

Write (θ; r, ϕ) for the coordinates on the copy of S 1 × D 2 (δ1) that we want<br />

to glue back. Then the contact form dθ + r 2 dϕ given on S 1 × D 2 (δ0) pulls back<br />

(along S 1 × ∂D 2 (δ1)) to<br />

d(nθ + qϕ) + r 2 d(mθ + pϕ).<br />

This form is defined on all of S 1 × (D 2 (δ1) − {0}), and to complete the proof it<br />

only remains to find a contact form on S 1 × D 2 (δ1) that coincides with this form<br />

near S 1 × ∂D 2 (δ1). It is at this point that we use an argument inspired by the<br />

Thurston-Winkelnkemper proof (but which goes back to Lutz).<br />

� �<br />

Lemma 3.8. Given a unimodular matrix<br />

n q<br />

m p<br />

, there is a contact form on<br />

S 1 × D 2 (δ) that coincides with (n + mr 2 )dθ + (q + pr 2 )dϕ near r = δ and with<br />

±dθ + r 2 dϕ near r = 0.<br />

Proof. We make the ansatz<br />

α = h1(r)dθ + h2(r)dϕ<br />

with smooth functions h1(r), h2(r). Then<br />

and<br />

dα = h ′ 1 dr ∧ dθ + h ′ 2 dr ∧ dϕ<br />

�<br />

�<br />

�<br />

α ∧ dα = �<br />

�<br />

h1 h2<br />

h ′ 1 h′ 2<br />

�<br />

�<br />

�<br />

� dθ ∧ dr ∧ dϕ.<br />

�<br />

So to satisfy the contact condition α ∧ dα �= 0 all we have to do is to find a<br />

parametrised curve<br />

r ↦−→ (h1(r), h2(r)), 0 ≤ r ≤ δ,<br />

in the plane satisfying the following conditions:<br />

1. h1(r) = ±1 and h2(r) = r 2 near r = 0,<br />

49


2. h1(r) = n + mr 2 and h2(r) = q + pr 2 near r = δ,<br />

3. (h1(r), h2(r)) is never parallel to (h ′ 1 (r), h′ 2 (r)).<br />

Since np−mq = ±1, the vector (m, p) is not a multiple of (n, q). Figure 12 shows<br />

possible solution curves for the two cases np − mq = ±1.<br />

−1<br />

h2<br />

(n,q)<br />

(n + m,q + p)<br />

h1<br />

h2<br />

Figure 12: Dehn surgery.<br />

(n + m,q + p)<br />

1<br />

(n,q)<br />

This completes the proof of the lemma and hence that of Theorem 3.7.<br />

Remark 3.9. On S 3 we have the standard contact forms α± = x dy − y dx ±<br />

(z dt−t dz) defining opposite orientations (cf. Remark 2.2). Performing the above<br />

surgery construction either on (S 3 , kerα+) or on (S 3 , kerα−) we obtain both<br />

positive and negative contact structures on any given M. The same is true for<br />

the Lutz construction that we study in the next two sections. Hence: A closed<br />

oriented 3–manifold admits both a positive and a negative contact structure in<br />

each homotopy class of tangent 2–plane fields.<br />

3.3 2–plane fields on 3–manifolds<br />

First we need the following well-known fact.<br />

Theorem 3.10. Every closed, orientable 3–manifold M is parallelisable.<br />

Remark. The most geometric proof of this theorem can be given based on a<br />

structure theorem of Hilden, Montesinos and Thickstun. This will be discussed<br />

in Section 3.5.2. Another proof can be found in [76]. Here we present the classical<br />

algebraic proof.<br />

50<br />

h1


Proof. The main point is to show the vanishing of the second Stiefel-Whitney<br />

class w2(M) = w2(TM) ∈ H 2 (M; Z2). Recall the following facts, which can<br />

be found in [14]; for the interpretation of Stiefel-Whitney classes as obstruction<br />

classes see also [95].<br />

There are Wu classes vi ∈ H i (M; Z2) defined by<br />

〈Sq i (u), [M]〉 = 〈vi ∪ u, [M]〉<br />

for all u ∈ H 3−i (M; Z2), where Sq denotes the Steenrod squaring operations.<br />

Since Sq i (u) = 0 for i > 3 − i, the only (potentially) non-zero Wu classes are<br />

v0 = 1 and v1. The Wu classes and the Stiefel-Whitney classes are related by<br />

wq = �<br />

j Sqq−j (vj). Hence v1 = Sq 0 (v1) = w1, which equals zero since M is<br />

orientable. We conclude w2 = 0.<br />

Let V2(R 3 ) = SO(3)/SO(1) = SO(3) be the Stiefel manifold of oriented,<br />

orthonormal 2–frames in R 3 . This is connected, so there exists a section over<br />

the 1–skeleton of M of the 2–frame bundle V2(TM) associated with TM (with a<br />

choice of Riemannian metric on M un<strong>der</strong>stood 4 ). The obstruction to extending<br />

this section over the 2–skeleton is equal to w2, which vanishes as we have just seen.<br />

The obstruction to extending the section over all of M lies in H 3 (M; π2(V2(R 3 ))),<br />

which is the zero group because of π2(SO(3)) = 0.<br />

We conclude that TM has a trivial 2–dimensional sub-bundle ε 2 . The com-<br />

plementary 1–dimensional bundle λ = TM/ε 2 is orientable and hence trivial<br />

since 0 = w1(TM) = w1(ε 2 ) + w1(λ) = w1(λ). Thus TM = ε 2 ⊕ λ is a trivial<br />

bundle.<br />

Fix an arbitrary Riemannian metric on M and a trivialisation of the unit<br />

tangent bundle STM ∼ = M × S 2 . This sets up a one-to-one correspondence<br />

between the following sets, where all maps, homotopies etc. are un<strong>der</strong>stood to be<br />

smooth.<br />

• Homotopy classes of unit vector fields X on M,<br />

• Homotopy classes of (co-)oriented 2–plane distributions ξ in TM,<br />

• Homotopy classes of maps f : M → S 2 .<br />

4 This is not necessary, of course. One may also work with arbitrary 2–frames without refer-<br />

ence to a metric. This does not affect the homotopical data.<br />

51


(I use the term ‘2–plane distribution’ synomymously with ‘2–dimensional sub-<br />

bundle of the tangent bundle’.) Let ξ1, ξ2 be two arbitrary 2–plane distributions<br />

(always un<strong>der</strong>stood to be cooriented). By elementary obstruction theory there is<br />

an obstruction<br />

d 2 (ξ1, ξ2) ∈ H 2 (M; π2(S 2 )) ∼ = H 2 (M; Z)<br />

for ξ1 to be homotopic to ξ2 over the 2–skeleton of M and, if d 2 (ξ1, ξ2) = 0 and<br />

after homotoping ξ1 to ξ2 over the 2–skeleton, an obstruction (which will depend,<br />

in general, on that first homotopy)<br />

d 3 (ξ1, ξ2) ∈ H 3 (M; π3(S 2 )) ∼ = H 3 (M; Z) ∼ = Z<br />

for ξ1 to be homotopic to ξ2 over all of M. (The identification of H 3 (M; Z) with Z<br />

is determined by the orientation of M.) However, rather than relying on general<br />

obstruction theory, we shall interpret d 2 and d 3 geometrically, which will later<br />

allow us to give a geometric proof that every homotopy class of 2–plane fields ξ<br />

on M contains a contact structure.<br />

The only fact that I want to quote here is that, by the Pontrjagin-Thom<br />

construction, homotopy classes of maps f : M → S 2 are in one-to-one correspon-<br />

dence with framed cobordism classes of framed (and oriented) links of 1–spheres<br />

in M. The general theory can be found in [14] and [77]; a beautiful and elemen-<br />

tary account is given in [94].<br />

For given f, the correspondence is defined by choosing a regular value p ∈ S 2<br />

for f and a positively oriented basis b of TpS 2 , and associating with it the oriented<br />

framed link (f −1 (p), f ∗ b), where f ∗ b is the pull-back of b un<strong>der</strong> the fibrewise<br />

bijective map Tf : T(f −1 (p)) ⊥ → TpS 2 . The orientation of f −1 (p) is the one<br />

which together with the frame f ∗ b gives the orientation of M.<br />

For a given framed link L the corresponding f is defined by projecting a<br />

(trivial) disc bundle neighbourhood L × D 2 of L in M onto the fibre D 2 ∼ =<br />

S 2 −p ∗ , where 0 is identified with p and p ∗ denotes the antipode of p, and sending<br />

M − (L × D 2 ) to p ∗ . Notice that the orientations of M and the components of L<br />

determine that of the fibre D 2 , and hence determine the map f.<br />

Before proceeding to define the obstruction classes d 2 and d 3 we make a<br />

short digression and discuss some topological background material which is fairly<br />

standard but not contained in our basic textbook references [14] and [77].<br />

52


3.3.1 Hopf’s Umkehrhomomorphismus<br />

If f : M m → N n is a continuous map between smooth manifolds, one can define<br />

a homomorphism ϕ : Hn−p(N) → Hm−p(M) on homology classes represented by<br />

submanifolds as follows. Given a homology class [L]N ∈ Hn−p(N) represented by<br />

a codimension p submanifold L, replace f by a smooth approximation transverse<br />

to L and define ϕ([L]N) = [f −1 (L)]M. This is essentially Hopf’s Umkehrhomo-<br />

morphismus [73], except that he worked with combinatorial manifolds of equal<br />

dimension and made no assumptions on the homology class. The following theo-<br />

rem, which in spirit is contained in [41], shows that ϕ is independent of choices (of<br />

submanifold L representing a class and smooth transverse approximation to f)<br />

and actually a homomorphism of intersection rings. This statement is not as well-<br />

known as it should be, and I only know of a proof in the literature for the special<br />

case where L is a point [60]. In [14] this map is called transfer map (more general<br />

transfer maps are discussed in [60]), but is only defined indirectly via Poincaré<br />

duality (though implicitly the statement of the following theorem is contained<br />

in [14], see for instance p. 377).<br />

Theorem 3.11. Let f : M m → N n be a smooth map between closed, oriented<br />

manifolds and L n−p ⊂ N n a closed, oriented submanifold of codimension p such<br />

that f is transverse to L. Write u ∈ H p (N) for the Poincaré dual of [L]N, that<br />

is, u∩[N] = [L]N. Then [f −1 (L)]M = f ∗ u∩[M]. In other words: If u is Poincaré<br />

dual to [L]N, then f ∗ u ∈ H p (M) is Poincaré dual to [f −1 (L)]M.<br />

Proof. Since f is transverse to L, the differential Tf induces a fibrewise isomor-<br />

phism between the normal bundles of f −1 (L) and L, and we find (closed) tubular<br />

neighbourhoods W → L and V = f −1 (W) → f −1 (L) (consi<strong>der</strong>ed as disc bun-<br />

dles) such that f : V → W is a fibrewise isomorphism. Write [V ]0 and [W]0 for<br />

the orientation classes in Hm(V, V − f −1 (L)) and Hn(W, W − L), respectively.<br />

We can identify these homology groups with Hm(V, ∂V ) and Hn(W, ∂W), respec-<br />

tively. Let τW ∈ H p (W, ∂W) and τV ∈ H p (V, ∂V ) be the Thom classes of these<br />

disc bundles defined by<br />

τW ∩ [W]0 = [L]N,<br />

τV ∩ [V ]0 = [f −1 (L)]M.<br />

Notice that f ∗ τW = τV since f : W → V is fibrewise isomorphic and the Thom<br />

class of an oriented disc bundle is the unique class whose restriction to each fibre<br />

53


is a positive generator of H p (D p , ∂D p ). Writing i: M → (M, M − f −1 (L)) and<br />

j : N → (N, N − L) for the inclusion maps we have<br />

[f −1 (L)]M = τV ∩ [V ]0<br />

= f ∗ τW ∩ [V ]0<br />

= f ∗ τW ∩ i∗[M],<br />

where we identify Hm(M, M −f −1 (L)) with Hm(V, V −f −1 (L)) un<strong>der</strong> the excision<br />

isomorphism. Then we have further<br />

[f −1 (L)]M = i ∗ f ∗ τW ∩ [M]<br />

= f ∗ j ∗ τW ∩ [M].<br />

So it remains to identify j ∗ τW as the Poincaré dual u of [L]N. Indeed,<br />

j ∗ τW ∩ [N] = τW ∩ j∗[N]<br />

= τW ∩ [W]0<br />

= [L]N,<br />

where we have used the excision isomorphism between the groups Hn(W, W −L)<br />

and Hn(N, N − L).<br />

3.3.2 Representing homology classes by submanifolds<br />

We now want to relate elements in H1(M; Z) to cobordism classes of links in M.<br />

Theorem 3.12. Let M be a closed, oriented 3–manifold. Any c ∈ H1(M; Z)<br />

is represented by an embedded, oriented link (of 1–spheres) Lc in M. Two links<br />

L0, L1 represent the same class [L0] = [L1] if and only if they are cobordant in M,<br />

that is, there is an embedded, oriented surface S in M × [0, 1] with<br />

where ⊔ denotes disjoint union.<br />

∂S = L1 ⊔ (−L0) ⊂ M × {1} ⊔ M × {0},<br />

Proof. Given c ∈ H1(M; Z), set u = PD(c) ∈ H 2 (M; Z), where PD denotes<br />

the Poincaré duality map from homology to cohomology. There is a well-known<br />

isomorphism<br />

H 2 (M; Z) ∼ = [M, K(Z, 2)] = [M, CP ∞ ],<br />

54


where brackets denote homotopy classes of maps (cf. [14, VII.12]). So u cor-<br />

responds to a homotopy class of maps [f]: M → CP ∞ such that f ∗ u0 = u,<br />

where u0 is the positive generator of H 2 (CP ∞ ) (that is, the one that pulls back<br />

to the Poincaré dual of [CP k−1 ] CP k un<strong>der</strong> the natural inclusion CP k ⊂ CP ∞ ).<br />

Since dimM = 3, any map f : M → CP ∞ is homotopic to a smooth map<br />

f1: M → CP 1 . Let p be a regular value of f1. Then<br />

PD(c) = u = f ∗ 1u0 = f ∗ 1PD[p] = PD[f −1<br />

1 (p)]<br />

by our discussion above, and hence c = [f −1<br />

link.<br />

1 (p)]. So Lc = f −1<br />

1<br />

(p) is the desired<br />

It is important to note that in spite of what we have just said it is not true that<br />

[M, CP ∞ ] = [M, CP 1 ], since a map F : M ×[0, 1] → CP ∞ with F(M × {0, 1}) ⊂<br />

CP 1 is not, in general, homotopic rel (M × {0, 1}) to a map into CP 1 . However,<br />

we do have [M, CP ∞ ] = [M, CP 2 ].<br />

If two links L0, L1 are cobordant in M, then clearly<br />

[L0] = [L1] ∈ H1(M × [0, 1]; Z) ∼ = H1(M; Z).<br />

For the converse, suppose we are given two links L0, L1 ⊂ M with [L0] = [L1].<br />

Choose arbitrary framings for these links and use this, as described at the be-<br />

ginning of this section, to define smooth maps f0, f1: M → S 2 with common<br />

regular value p ∈ S 2 such that f −1<br />

i (p) = Li, i = 0, 1. Now identify S 2 with the<br />

standardly embedded CP 1 ⊂ CP 2 . Let P ⊂ CP 2 be a second copy of CP 1 , em-<br />

bedded in such a way that [P] CP 2 = [CP 1 ] CP 2 and P intersects CP 1 transversely<br />

in p only. This is possible since CP 1 ⊂ CP 2 has self-intersection one. Then<br />

the maps f0, f1, regarded as maps into CP 2 , are transverse to P and we have<br />

f −1<br />

i (P) = Li, i = 0, 1. Hence<br />

f ∗ i u0 = f ∗ i (PD[P] CP 2) = PD[f −1<br />

i (P)]M<br />

= PD[Li]M<br />

is the same for i = 0 or 1, and from the identification<br />

[M, CP 2 ]<br />

∼=<br />

−→ H 2 (M, Z)<br />

[f] ↦−→ f ∗ u0<br />

we conclude that f0 and f1 are homotopic as maps into CP 2 .<br />

55


Let F : M × [0, 1] → CP 2 be a homotopy between f0 and f1, which we may<br />

assume to be constant near 0 and 1. This F can be smoothly approximated by a<br />

map F ′ : M × [0, 1] → CP 2 which is transverse to P and coincides with F near<br />

M ×0 and M ×1 (since there the transversality condition was already satisfied).<br />

In particular, F ′ is still a homotopy between f0 and f1, and S = (F ′ ) −1 (P) is a<br />

surface with the desired property ∂S = L1 ⊔ (−L0).<br />

Notice that in the course of this proof we have observed that cobordism classes<br />

of links in M (equivalently, classes in H1(M; Z)) correspond to homotopy classes<br />

of maps M → CP 2 , whereas framed cobordism classes of framed links correspond<br />

to homotopy classes of maps M → CP 1 .<br />

By forming the connected sum of the components of a link representing a<br />

certain class in H1(M; Z), one may actually always represent such a class by a<br />

link with only one component, that is, a knot.<br />

3.3.3 Framed cobordisms<br />

We have seen that if L1, L2 ⊂ M are links with [L1] = [L2] ∈ H1(M; Z), then<br />

L1 and L2 are cobordant in M. In general, however, a given framing on L1 and<br />

L2 does not extend over the cobordism. The following observation will be useful<br />

later on.<br />

Write (S 1 , n) for a contractible loop in M with framing n ∈ Z (by which<br />

we mean that S 1 and a second copy of S 1 obtained by pushing it away in the<br />

direction of one of the vectors in the frame have linking number n). When writing<br />

L = L ′ ⊔ (S 1 , n) it is un<strong>der</strong>stood that (S 1 , n) is not linked with any component<br />

of L ′ .<br />

Suppose we have two framed links L0, L1 ⊂ M with [L0] = [L1]. Let S ⊂<br />

M × [0, 1] be an embedded surface with<br />

∂S = L1 ⊔ (−L0) ⊂ M × {1} ⊔ M × {0}.<br />

With D 2 a small disc embedded in S, the framing of L1 and L2 in M extends<br />

to a framing of S − D 2 in M × [0, 1] (since S − D 2 deformation retracts to a 1–<br />

dimensional complex containing L0 and L1, and over such a complex an orientable<br />

2–plane bundle is trivial). Now we embed a cylin<strong>der</strong> S 1 ×[0, 1] in M ×[0, 1] such<br />

that<br />

S 1 × [0, 1] ∩ M × {0} = ∅,<br />

56


and<br />

S 1 × [0, 1] ∩ M × {1} = S 1 × {1},<br />

S 1 × [0, 1] ∩ (S − D 2 ) = S 1 × {0} = ∂D 2 ,<br />

see Figure 13. This shows that L0 is framed cobordant in M to L1 ⊔ (S 1 , n) for<br />

suitable n ∈ Z.<br />

L0 ⊂ M × {0}<br />

D 2<br />

S<br />

S 1 × [0,1]<br />

L1 ⊂ M × {1}<br />

Figure 13: The framed cobordism between L0 and L1 ⊔ (S 1 , n).<br />

3.3.4 Definition of the obstruction classes<br />

We are now in a position to define the obstruction classes d 2 and d 3 . With a<br />

choice of Riemannian metric on M and a trivialisation of STM un<strong>der</strong>stood, a<br />

2–plane distribution ξ on M defines a map fξ : M → S 2 and hence an oriented<br />

framed link Lξ as described above. Let [Lξ] ∈ H1(M; Z) be the homology class<br />

represented by Lξ. This only depends on the homotopy class of ξ, since un<strong>der</strong><br />

homotopies of ξ or choice of different regular values of fξ the cobordism class of<br />

Lξ remains invariant. We define<br />

d 2 (ξ1, ξ2) = PD[Lξ1 ] − PD[Lξ2 ].<br />

With this definition d 2 is clearly additive, that is,<br />

d 2 (ξ1, ξ2) + d 2 (ξ2, ξ3) = d 2 (ξ1, ξ3).<br />

The following lemma shows that d 2 is indeed the desired obstruction class.<br />

57


Lemma 3.13. The 2–plane distributions ξ1 and ξ2 are homotopic over the 2–<br />

skeleton M (2) of M if and only if d 2 (ξ1, ξ2) = 0.<br />

Proof. Suppose d2 (ξ1, ξ2) = 0, that is, [Lξ1 ] = [Lξ2 ]. By Theorem 3.12 we find a<br />

surface S in M × [0, 1] with<br />

∂S = Lξ2 ⊔ (−Lξ1 ) ⊂ M × {1} ⊔ M × {0}.<br />

From the discussion on framed cobordism above we know that for suitable n ∈ Z<br />

we find a framed surface S ′ in M × [0, 1] such that<br />

as framed manifolds.<br />

∂S ′ = � Lξ2 ⊔ (S1 , n) � ⊔ (−Lξ1 ) ⊂ M × {1} ⊔ M × {0}<br />

Hence ξ1 is homotopic to a 2–plane distribution ξ ′ 1 such that L ξ ′ 1<br />

and Lξ2 differ<br />

only by one contractible framed loop (not linked with any other component).<br />

Then the corresponding maps f ′ 1 , f2 differ only in a neighbourhood of this loop,<br />

which is contained in a 3–ball, so f ′ 1 and f2 (and hence ξ ′ 1 and ξ2) agree over the<br />

2–skeleton.<br />

Conversely, if ξ1 and ξ2 are homotopic over M (2) , we may assume ξ1 = ξ2 on<br />

M − D3 for some embedded 3–disc D3 ⊂ M without changing [Lξ1 ] and [Lξ2 ].<br />

Now [Lξ1 ] = [Lξ2 ] follows from H1(D3 , S2 ) = 0.<br />

Remark 3.14. By [99, § 37] the obstruction to homotopy between ξ and ξ0<br />

(corresponding to the constant map fξ0 : M → S2 ) over the 2–skeleton of M is<br />

given by f ∗ ξ u0, where u0 is the positive generator of H 2 (S 2 ; Z). So u0 = PD[p]<br />

for any p ∈ S 2 , and taking p to be a regular value of fξ we have<br />

f ∗ ξ u0 = f ∗ −1<br />

ξ PD[p] = PD[fξ (p)]<br />

= PD[Lξ] = d 2 (ξ, ξ0).<br />

This gives an alternative way to see that our geometric definition of d 2 does<br />

indeed coincide with the class defined by classical obstruction theory.<br />

Now suppose d 2 (ξ1, ξ2) = 0. We may then assume that ξ1 = ξ2 on M−int(D 3 ),<br />

and we define d 3 (ξ1, ξ2) to be the Hopf invariant H(f) of the map f : S 3 → S 2<br />

defined as f1 ◦ π+ on the upper hemisphere and f2 ◦ π− on the lower hemisphere,<br />

where π+, π− are the orthogonal projections of the upper resp. lower hemisphere<br />

58


onto the equatorial disc, which we identify with D 3 ⊂ M. Here, given an orien-<br />

tation of M, we orient S 3 in such a way that π+ is orientation-preserving and<br />

π− orientation-reversing; the orientation of S 2 is inessential for the computation<br />

of H(f). Recall that H(f) is defined as the linking number of the preimages of<br />

two distinct regular values of a smooth map homotopic to f. Since the Hopf in-<br />

variant classifies homotopy classes of maps S 3 → S 2 (it is in fact an isomorphism<br />

π3(S 2 ) → Z), this is a suitable definition for the obstruction class d 3 . Moreover,<br />

the homomorphism property of H(f) and the way addition in π3(S 2 ) is defined<br />

entail the additivity of d 3 analogous to that of d 2 .<br />

For M = S 3 there is another way to interpret d 3 . Oriented 2–plane distrib-<br />

utions on M correspond to sections of the bundle associated to TM with fibre<br />

SO(3)/U(1), hence to maps M → SO(3)/U(1) ∼ = S 2 since TM is trivial. Simi-<br />

larly, almost complex structures on D 4 correspond to maps D 4 → SO(4)/U(2) ∼ =<br />

SO(3)/U(1) (cf. [61] for this isomorphism). A cooriented 2–plane distribution on<br />

M can be interpreted as a triple (X, ξ, J), where X is a vector field transverse<br />

to ξ defining the coorientation, and J a complex structure on ξ defining the ori-<br />

entation. Such a triple is called an almost contact structure. (This notion<br />

generalises to higher (odd) dimensions, and by Remark 2.3 every cooriented con-<br />

tact structure induces an almost contact structure, and in fact a unique one up<br />

to homotopy as follows from the result cited in that remark.) Given an almost<br />

contact structure (X, ξ, J) on S 3 , one defines an almost complex structure � J on<br />

TD 4 |S 3 by � J|ξ = J and � JN = X, where N denotes the outward normal vector<br />

field. So there is a canonical way to identify homotopy classes of almost con-<br />

tact structures on S 3 with elements of π3(SO(3)/U(1)) ∼ = Z such that the value<br />

zero corresponds to the almost contact structure that extends as almost complex<br />

structure over D 4 .<br />

3.4 Let’s Twist Again<br />

Consi<strong>der</strong> a 3–manifold M with cooriented contact structure ξ and an oriented 1–<br />

sphere K ⊂ M embedded transversely to ξ such that the positive orientation of K<br />

coincides with the positive coorientation of ξ. Then in suitable local coordinates<br />

we can identify K with S 1 × {0} ⊂ S 1 × D 2 such that ξ = ker(dθ + r 2 dϕ) and<br />

∂θ corresponds to the positive orientation of K (see Example 2.33). Strictly<br />

speaking, if, as we shall always assume, S 1 is parametrised by 0 ≤ θ ≤ 2π, then<br />

this formula for ξ holds on S 1 × D 2 (δ) for some, possibly small, δ > 0. However,<br />

59


to simplify notation we usually work with S 1 × D 2 as local model.<br />

We say that ξ ′ is obtained from ξ by a Lutz twist along K and write ξ ′ = ξ K<br />

if on S 1 × D 2 the new contact structure ξ ′ is defined by<br />

ξ ′ = ker(h1(r)dθ + h2(r)dϕ)<br />

with (h1(r), h2(r)) as in Figure 14, and ξ ′ coincides with ξ outside S 1 × D 2 .<br />

h2<br />

r = r0<br />

−1 1<br />

Figure 14: Lutz twist.<br />

More precisely, (h1(r), h2(r)) is required to satisfy the conditions<br />

1. h1(r) = −1 and h2(r) = −r 2 near r = 0,<br />

2. h1(r) = 1 and h2(r) = r 2 near r = 1,<br />

3. (h1(r), h2(r)) is never parallel to (h ′ 1 (r), h′ 2 (r)).<br />

This is the same as applying the construction � of Section � 3.2 to the topologically<br />

trivial Dehn surgery given by the matrix<br />

−1<br />

0<br />

0<br />

−1<br />

.<br />

It will be useful later on to un<strong>der</strong>stand more precisely the behaviour of the<br />

map fξ ′ : S3 → S 2 . For the definition of this map we assume – this assump-<br />

tion will be justified below – that on S 1 × D 2 the map fξ was defined in terms<br />

60<br />

h1


of the standard metric dθ 2 + du 2 + dv 2 (with u, v cartesian coordinates on D 2<br />

corresponding to the polar coordinates r, ϕ) and the trivialisation ∂θ, ∂u, ∂v of<br />

T(S 1 × D 2 ). Since ξ ′ is spanned by ∂r and h2(r)∂θ − h1(r)∂ϕ (resp. ∂u, ∂v for<br />

r = 0), a vector positively orthogonal to ξ ′ is given by<br />

h1(r)∂θ + h2(r)∂ϕ,<br />

which makes sense even for r = 0. Observe that the ratio h1(r)/h2(r) (for<br />

h2(r) �= 0) is a strictly monotone decreasing function since by the third condition<br />

above we have<br />

(h1/h2) ′ = (h ′ 1h2 − h1h ′ 2)/h 2 2 < 0.<br />

This implies that any value on S 2 other than (1, 0, 0) (corresponding to ∂θ) is<br />

regular for the map fξ ′; the value (1, 0, 0) is attained along the torus {r = r0},<br />

with r0 > 0 determined by h2(r0) = 0, and hence not regular.<br />

If S 1 × D 2 is endowed with the orientation defined by the volume form dθ ∧<br />

r dr ∧ dϕ = dθ ∧ du ∧ dv (so that ξ and ξ ′ are positive contact structures) and<br />

S 2 ⊂ R 3 is given its ‘usual’ orientation defined by the volume form x dy ∧ dz +<br />

y dz ∧ dx + z dx ∧ dy, then<br />

f −1<br />

ξ ′ (−1, 0, 0) = S 1 × {0}<br />

with orientation given by −∂θ, since fξ ′ maps the slices {θ} ×D2 (r0) orientation-<br />

reversingly onto S 2 .<br />

More generally, for any p ∈ S 2 − {(1, 0, 0)} the preimage f −1<br />

ξ ′ (p) (inside the<br />

domain {(θ, r, ϕ): h2(r) < 0} = {r = r0}) is a circle S 1 × {u}, u ∈ D 2 , with<br />

orientation given by −∂θ.<br />

We are now ready to show how to construct a contact structure on M in<br />

any given homotopy class of 2–plane distributions by starting with an arbitrary<br />

contact structure and performing suitable Lutz twists. First we deal with homo-<br />

topy over the 2–skeleton. One way to proceed would be to prove directly, with<br />

notation as above, that d 2 (ξ K , ξ) = −PD[K]. However, it is somewhat easier<br />

to compute d 2 (ξ K , ξ) in the case where ξ is a trivial 2–plane bundle and the<br />

trivialisation of STM is adapted to ξ. Since I would anyway like to present an<br />

alternative argument for computing the effect of a Lutz twist on the Euler class<br />

of the contact structure, and thus relate d 2 (ξ1, ξ2) with the Euler classes of ξ1<br />

and ξ2, it seems opportune to do this first and use it to show the existence of<br />

61


a contact structure with Euler class zero. In the next section we shall actually<br />

discuss a direct geometric proof, due to Gonzalo, of the existence of a contact<br />

structure with Euler class zero.<br />

Recall that the Euler class e(ξ) ∈ H 2 (B; Z) of a 2–plane bundle over a complex<br />

B (of arbitrary dimension) is the obstruction to finding a nowhere zero section<br />

of ξ over the 2–skeleton of B. Since πi(S 1 ) = 0 for i ≥ 2, all higher obstruction<br />

groups H i+1 (B; πi(S 1 )) are trivial, so a 2–dimensional orientable bundle ξ is<br />

trivial if and only if e(ξ) = 0, no matter what the dimension of B.<br />

Now let ξ be an arbitrary cooriented 2–plane distribution on an oriented 3–<br />

manifold M. Then TM ∼ = ξ ⊕ ε 1 , where ε 1 denotes a trivial line bundle. Hence<br />

w2(ξ) = w2(ξ ⊕ ε 1 ) = w2(TM) = 0, and since w2(ξ) is the mod 2 reduction of<br />

e(ξ) we infer that e(ξ) has to be even.<br />

Proposition 3.15. For any even element e ∈ H 2 (M; Z) there is a contact struc-<br />

ture ξ on M with e(ξ) = e.<br />

Proof. Start with an arbitrary contact structure ξ0 on M with e(ξ0) = e0 (which<br />

we know to be even). Given any even e1 ∈ H 2 (M; Z), represent the Poincaré dual<br />

of (e0 − e1)/2 by a collection of embedded oriented circles positively transverse<br />

to ξ0. (Here by (e0 − e1)/2 I mean some class whose double equals e0 − e1; in<br />

the presence of 2–torsion there is of course a choice of such classes.) Choose<br />

a section of ξ0 transverse to the zero section of ξ0, that is, a vector field in<br />

ξ0 with generic zeros. We may assume that there are no zeros on the curves<br />

representing PD −1 (e0 − e1)/2. Now perform a Lutz twist as described above<br />

along these curves and call ξ1 the resulting contact structure. It is easy to see<br />

that in the local model for the Lutz twist a constant vector field tangent to ξ0<br />

along ∂(S 1 ×D 2 (r0)) extends to a vector field tangent to ξ1 over S 1 ×D 2 (r0) with<br />

zeros of index +2 along S 1 × {0} (Figure 15). So the vector field in ξ0 extends<br />

to a vector field in ξ1 with new zeros of index +2 along the curves representing<br />

PD −1 (e1 − e0)/2 (notice that a Lutz twist along a positively transverse knot K<br />

turns K into a negatively transverse knot). Since the self-intersection class of M<br />

in the total space of a vector bundle is Poincaré dual to the Euler class of that<br />

bundle, this proves e(ξ1) = e(ξ0) + e1 − e0 = e1.<br />

We now fix a contact structure ξ0 on M with e(ξ0) = 0 and give M the ori-<br />

entation induced by ξ0 (i.e. the one for which ξ0 is a positive contact structure).<br />

Moreover, we fix a Riemannian metric on M and define X0 as the vector field<br />

62


2<br />

1<br />

2<br />

1<br />

1<br />

1<br />

2<br />

2<br />

1<br />

2 2<br />

Figure 15: Effect of Lutz twist on Euler class.<br />

positively orthonormal to ξ0. Since ξ0 is a trivial plane bundle, X0 extends to an<br />

orthonormal frame X0, X1, X2, hence a trivialisation of STM, with X1, X2 tan-<br />

gent to ξ0 and defining the orientation of ξ0. With these choices, ξ0 corresponds<br />

to the constant map fξ0 : M → (1, 0, 0) ∈ S2 .<br />

Proposition 3.16. Let K ⊂ M be an embedded, oriented circle positively trans-<br />

verse to ξ0. Then d 2 (ξ K 0 , ξ0) = −PD[K].<br />

Proof. Identify a tubular neighbourhood of K ⊂ M with S 1 × D 2 with framing<br />

defined by X1, and ξ0 given in this neighbourhood as the kernel of dθ + r 2 dϕ =<br />

dθ+u dv−v du. We may then change the trivialisation X0, X1, X2 by a homotopy,<br />

fixed outside S 1 × D 2 , such that X0 = ∂θ, X1 = ∂u and X2 = ∂v near K; this<br />

does not change the homotopical data of 2–plane distributions computed via the<br />

Pontrjagin-Thom construction. Then fξ0 is no longer constant, but its image still<br />

does not contain the point (−1, 0, 0).<br />

Now perform a Lutz twist along K × {0}. Our discussion at the beginning<br />

of this section shows that (−1, 0, 0) is a regular value of the map fξ : M → S 2<br />

associated with ξ = ξK −1<br />

0 and fξ (−1, 0, 0) = −K. Hence, by definition of the<br />

obstruction class d2 we have d2 (ξK 0 , ξ0) = −PD[K].<br />

Proof of Theorem 3.1. Let η be a 2–plane distribution on M and ξ0 the contact<br />

structure on M with e(ξ0) = 0 that we fixed earlier on. According to our discus-<br />

sion in Section 3.3.2 and Theorem 2.44, we can find an oriented knot K positively<br />

transverse to ξ0 with −PD[K] = d 2 (η, ξ0). Then d 2 (η, ξ0) = d 2 (ξ K 0 , ξ0) by the<br />

preceding proposition, and therefore d 2 (ξ K 0<br />

63<br />

2<br />

1<br />

, η) = 0.<br />

1<br />

2<br />

1<br />

2<br />

1


We may then assume that η = ξ K 0 on M − D3 , where we choose D 3 so<br />

small that ξ K 0 is in Darboux normal form there (and identical with ξ0). By<br />

Proposition 3.4 we can find a link K ′ in D3 transverse to ξK 0<br />

number l(K ′ ) equal to d3 (η, ξK 0 ).<br />

with self-linking<br />

Now perform a Lutz twist of ξ K 0 along each component of K′ and let ξ be the<br />

resulting contact structure. Since this does not change ξ K 0<br />

of M, we still have d 2 (ξ, η) = 0.<br />

over the 2–skeleton<br />

Observe that f ξ K 0 | D 3 does not contain the point (−1, 0, 0) ∈ S 2 , and – since<br />

f ξ K 0 (D 3 ) is compact – there is a whole neighbourhood U ⊂ S 2 of (−1, 0, 0) not<br />

contained in f ξ K 0 (D 3 ). Let f : S 3 → S 2 be the map used to compute d 3 (ξ, ξ K 0 ),<br />

that is, f coincides on the upper hemisphere with fξ| D 3 and on the lower hemi-<br />

sphere with f ξ K 0 | D 3. By the discussion in Section 3.3, the preimage f −1 (u) of any<br />

u ∈ U − {(−1, 0, 0)} will be a push-off of −K ′ determined by the trivialisation of<br />

ξ K 0 | D 3 = ξ0| D 3. So the linking number of f −1 (u) with f −1 (−1, 0, 0), which is by<br />

definition the Hopf invariant H(f) = d 3 (ξ, ξ K 0 ), will be equal to l(K′ ). By our<br />

choice of K ′ and the additivity of d 3 this implies d 3 (ξ, η) = 0. So ξ is a contact<br />

structure that is homotopic to η as a 2–plane distribution.<br />

3.5 Other existence proofs<br />

Here I briefly summarise the other known existence proofs for contact structures<br />

on 3–manifolds, mostly by pointing to the relevant literature. In spirit, most of<br />

these proofs are similar to the one by Lutz-Martinet: start with a structure theo-<br />

rem for 3–manifolds and show that the topological construction can be performed<br />

compatibly with a contact structure.<br />

3.5.1 Open books<br />

According to a theorem of Alexan<strong>der</strong> [5], cf. [97], every closed, orientable 3–<br />

manifold M admits an open book decomposition. This means that there is<br />

a link L ⊂ M, called the binding, and a fibration f : M − L → S 1 , whose fibres<br />

are called the pages, see Figure 16. It may be assumed that L has a tubular<br />

neighbourhood L × D 2 such that f restricted to L × (D 2 − {0}) is given by<br />

f(θ, r, ϕ) = ϕ, where θ is the coordinate along L and (r, ϕ) are polar coordinates<br />

on D 2 .<br />

At the cost of raising the genus of the pages, one may decrease the number<br />

of components of L, and in particular one may always assume L to be a knot.<br />

64


S 1<br />

f −1 (ϕ)<br />

L<br />

Figure 16: An open book near the binding.<br />

Another way to think of such an open book is as follows. Start with a surface<br />

Σ with boundary ∂Σ = K ∼ = S 1 and a self-diffeomorphism h of Σ with h = id<br />

near K. Form the mapping torus Th = Σh = Σ × [0, 2π]/∼, where ‘∼’ denotes<br />

the identification (p, 2π) ∼ (h(p), 0). Define a 3–manifold M by<br />

M = Th ∪ K×S 1 (K × D 2 ).<br />

This M carries by construction the structure of an open book with binding K<br />

and pages diffeomorphic to Σ.<br />

Here is a slight variation on a simple argument of Thurston and Winkelnkem-<br />

per [101] for producing a contact structure on such an open book (and hence on<br />

any closed, orientable 3–manifold):<br />

Start with a 1–form β0 on Σ with β0 = e t dθ near ∂Σ = K, where θ denotes<br />

the coordinate along K and t is a collar parameter with K = {t = 0} and t < 0 in<br />

the interior of Σ. Then dβ0 integrates to 2π over Σ by Stokes’s theorem. Given<br />

any area form ω on Σ (with total area equal to 2π) satisfying ω = e t dt ∧ dθ<br />

near K, the 2–form ω −dβ0 is, by de Rham’s theorem, an exact 1–form, say dβ1,<br />

where we may assume β1 ≡ 0 near K.<br />

Set β = β0 + β1. Then dβ = ω is an area form (of total area 2π) on Σ and<br />

β = e t dθ near K. The set of 1–forms satisfying these two properties is a convex<br />

set, so we can find a 1–form (still denoted β) on Th which has these properties<br />

when restricted to the fibre over any ϕ ∈ S 1 = [0, 2π]/0∼2π. We may (and shall)<br />

65


equire that β = e t dθ near ∂Th.<br />

Now a contact form α on Th is found by setting α = β+C dϕ for a sufficiently<br />

large constant C ∈ R + , so that in<br />

α ∧ dα = (β + C dϕ) ∧ dβ<br />

the non-zero term dϕ ∧ dβ = dϕ ∧ ω dominates. This contact form can be<br />

extended to all of M by making the ansatz α = h1(r)dθ + h2(r)dϕ on K × D 2 ,<br />

as described in our discussion of the Lutz twist. The boundary conditions in the<br />

present situation are, say,<br />

1. h1(r) = 2 and h2(r) = r 2 near r = 0,<br />

2. h1(r) = e 1−r and h2(r) = C near r = 1.<br />

Observe that subject to these boundary conditions a curve (h1(r), h2(r)) can<br />

be found that does not pass the h2–axis (i.e. with h1(r) never being equal to<br />

zero). In the 3–dimensional setting this is not essential (and the Thurston-<br />

Winkelnkemper ansatz lacked that feature), but it is crucial when one tries to<br />

generalise this construction to higher dimensions. This has recently been worked<br />

out by Giroux and J.-P. Mohsen [57]. This, however, is only the easy part of<br />

their work. Rather strikingly, they have shown that a converse of this result<br />

holds: Given a compact contact manifold of arbitrary dimension, it admits an<br />

open book decomposition that is adapted to the contact structure in the way<br />

described above. Full details have not been published at the time of writing, but<br />

see Giroux’s talk [56] at the ICM 2002.<br />

3.5.2 Branched covers<br />

A theorem of Hilden, Montesinos and Thickstun [63] states that every closed,<br />

orientable 3–manifold M admits a branched covering π: M → S 3 such that<br />

the upstairs branch set is a simple closed curve that bounds an embedded disc.<br />

(Moreover, the cover can be chosen 3–fold and simple, i.e. the monodromy repre-<br />

sentation of π1(S 3 −K), where K is the branching set downstairs (a knot in S 3 ),<br />

represents the meridian of K by a transposition in the symmetric group S3. This,<br />

however, is not relevant for our discussion.)<br />

It follows immediately, as announced in Section 3.3, that every closed, ori-<br />

entable 3–manifold is parallelisable: First of all, S 3 is parallelisable since it car-<br />

ries a Lie group structure (as the unit quaternions, for instance). Given M and<br />

66


a branched covering π: M → S 3 as above, there is a 3–ball D 3 ⊂ M containing<br />

the upstairs branch set. Outside of D 3 , the covering π is unbranched, so the<br />

3–frame on S 3 can be lifted to a frame on M −D 3 . The bundle TM| D 3 is trivial,<br />

so the frame defined along ∂D 3 defines an element of SO(3) (cf. the footnote in<br />

the proof of Theorem 3.10). Since π2(SO(3)) = 0, this frame extends over D 3 .<br />

In [59], Gonzalo uses this theorem to construct a contact structure on every<br />

closed, orientable 3–manifold M, in fact one with zero Euler class: Away from<br />

the branching set one can lift the standard contact structure from S 3 (which<br />

has Euler class zero: a trivialisation is given by two of the three (quaternionic)<br />

Hopf vector fields). A careful analysis of the branched covering map near the<br />

branching set then shows how to extend this contact structure over M (while<br />

keeping it trivial as 2–plane bundle).<br />

A branched covering construction for higher-dimensional contact manifolds is<br />

discussed in [43].<br />

3.5.3 . . . and more<br />

The work of Giroux [52], in which he initiated the study of convex surfaces in<br />

contact 3–manifolds, also contains a proof of Martinet’s theorem.<br />

An entirely different proof, due to S. Altschuler [4], finds contact structures<br />

from solutions to a certain parabolic differential equation for 1–forms on 3–<br />

manifolds. Some of these ideas have entered into the more far-reaching work<br />

of Eliashberg and Thurston on so-called ‘confoliations’ [32], that is, 1–forms sat-<br />

isfying α ∧ dα ≥ 0.<br />

3.6 Tight and overtwisted<br />

The title of this section describes the fundamental dichotomy of contact structures<br />

in dimension 3 that has proved seminal for the development of the field.<br />

In or<strong>der</strong> to motivate the notion of an overtwisted contact structure, as intro-<br />

duced by Eliashberg [21], we consi<strong>der</strong> a ‘full’ Lutz twist as follows. Let (M, ξ) be<br />

a contact 3–manifold and K ⊂ M a knot transverse to ξ. As before, identify K<br />

with S 1 × {0} ⊂ S 1 × D 2 ⊂ M such that ξ = ker(dθ + r 2 dϕ) on S 1 × D 2 . Now<br />

define a new contact structure ξ ′ as in Section 3.4, with (h1(r), h2(r)) now as in<br />

Figure 17, that is, the boundary conditions are now<br />

h1(r) = 1 and h2(r) = r 2 for r ∈ [0, ε] ∪ [1 − ε, 1]<br />

67


for some small ε > 0.<br />

h2<br />

Figure 17: A full Lutz twist.<br />

Lemma 3.17. A full Lutz twist does not change the homotopy class of ξ as a<br />

2–plane field.<br />

Proof. Let (ht 1 (r), ht2 (r)), r, t ∈ [0, 1], be a homotopy of paths such that<br />

1. h 0 1 ≡ 1, h0 2 (r) = r2 ,<br />

2. h 1 i ≡ hi, i = 1, 2,<br />

3. h t i (r) = hi(r) for r ∈ [0, ε] ∪ [1 − ε, 1].<br />

Let χ : [0, 1] → R be a smooth function which is identically zero near r = 0 and<br />

r = 1 and χ(r) > 0 for r ∈ [ε, 1 − ε]. Then<br />

αt = t(1 − t)χ(r)dr + h t 1(r)dθ + h t 2(r)dϕ<br />

is a homotopy from α0 = dθ +r 2 dϕ to α1 = h1(r)dθ +h2(r)dϕ through non-zero<br />

1–forms. This homotopy stays fixed near r = 1, and so it defines a homotopy<br />

between ξ and ξ ′ as 2–plane fields.<br />

68<br />

1<br />

h1


Let r0 be the smaller of the two positive radii with h2(r0) = 0 and consi<strong>der</strong><br />

the embedding<br />

φ : D 2 (r0) −→ S 1 × D 2<br />

(r, ϕ) ↦−→ (θ(r), r, ϕ),<br />

where θ(r) is a smooth function with θ(r0) = 0, θ(r) > 0 for 0 ≤ r < r0, and<br />

θ ′ (r) = 0 only for r = 0. We may require in addition that θ(r) = θ(0) − r 2 near<br />

r = 0. Then<br />

φ ∗ (h1(r)dθ + h2(r)dϕ) = h1(r)θ ′ (r)dr + h2(r)dϕ<br />

is a differential 1–form on D 2 (r0) which vanishes only for r = 0, and along<br />

∂D 2 (r0) the vector field ∂ϕ tangent to the boundary lies in the kernel of this 1–<br />

form, see Figure 18. In other words, the contact planes ker(h1(r)dθ + h2(r)dϕ)<br />

intersected with the tangent planes to the embedded disc φ(D 2 (r0)) induce a<br />

singular 1–dimensional foliation on this disc with the boundary of this disc as<br />

closed leaf and precisely one singular point in the interior of the disc (Figure 19;<br />

notice that the leaves of this foliation are the integral curves of the vector field<br />

h1(r)θ ′ (r)∂ϕ − h2(r)∂r). Such a disc is called an overtwisted disc.<br />

S 1<br />

r0<br />

Figure 18: An overtwisted disc.<br />

φ(D 2 (r0))<br />

A contact structure ξ on a 3–manifold M is called overtwisted if (M, ξ)<br />

contains an embedded overtwisted disc. In or<strong>der</strong> to justify this terminology,<br />

69<br />

ξ


Figure 19: Characteristic foliation on an overtwisted disc.<br />

observe that in the radially symmetric standard contact structure of Example 2.7,<br />

the angle by which the contact planes turn approaches π/2 asymptotically as r<br />

goes to infinity. By contrast, any contact manifold which has been constructed<br />

using at least one (simple) Lutz twist contains a similar cylindrical region where<br />

the contact planes twist by more than π in radial direction (at the smallest<br />

positive radius r0 with h2(r0) = 0 the twisting angle has reached π).<br />

We have shown the following:<br />

Proposition 3.18. Let ξ be a contact structure on M. By a full Lutz twist along<br />

any transversely embedded circle one obtains an overtwisted contact structure ξ ′<br />

that is homotopic to ξ as a 2–plane distribution. �<br />

Together with the theorem of Lutz and Martinet we find that M contains an<br />

overtwisted contact structure in every homotopy class of 2–plane distributions.<br />

In fact, Eliashberg [21] has proved the following much stronger theorem.<br />

Theorem 3.19 (Eliashberg). On a closed, orientable 3–manifold there is a one-<br />

to-one correspondence between homotopy classes of overtwisted contact structures<br />

and homotopy classes of 2–plane distributions.<br />

This means that two overtwisted contact structures which are homotopic as<br />

2–plane fields are actually homotopic as contact structures and hence isotopic by<br />

Gray’s stability theorem.<br />

Thus, it ‘only’ remains to classify contact structures that are not overtwisted.<br />

In [24] Eliashberg defined tight contact structures on a 3–manifold M as contact<br />

structures ξ for which there is no embedded disc D ⊂ M such that Dξ contains<br />

70


a limit cycle. So, by definition, overtwisted contact structures are not tight. In<br />

that same paper, as mentioned above in Section 2.4.5, Eliashberg goes on to show<br />

the converse with the help of the Elimination Lemma: non-overtwisted contact<br />

structures are tight.<br />

There are various ways to detect whether a contact structure is tight. His-<br />

torically the first proof that a certain contact structure is tight is due to D. Ben-<br />

nequin [9, Cor. 2, p. 150]:<br />

Theorem 3.20 (Bennequin). The standard contact structure ξ0 on S 3 is tight.<br />

The steps of the proof are as follows: (i) First, Bennequin shows that if γ0 is<br />

a transverse knot in (S 3 , ξ0) with Seifert surface Σ, then the self-linking number<br />

of γ satisfies the inequality<br />

l(γ0) ≤ −χ(Σ).<br />

(ii) Second, he introduces an invariant for Legendrian knots; nowadays this<br />

is called the Thurston-Bennequin invariant: Let γ be a Legendrian knot in<br />

(S 3 , ξ0). Take a vector field X along γ transverse to ξ0, and let γ ′ be the push-<br />

off of γ in the direction of X. Then the Thurston-Bennequin invariant tb(γ) is<br />

defined to be the linking number of γ and γ ′ . (This invariant has an extension<br />

to homologically trivial Legendrian knots in arbitrary contact 3–manifolds.)<br />

(iii) By pushing γ in the direction of ±X, one obtains transverse curves γ ±<br />

(either of which is a candidate for γ ′ in (ii)). One of these curves is positively<br />

transverse, the other negatively transverse to ξ0. The self-linking number of γ ± is<br />

related to the Thurston-Bennequin invariant and a further invariant (the rotation<br />

number) of γ. The equation relating these three invariants implies tb(γ) ≤ −χ(Σ),<br />

where Σ again denotes a Seifert surface for γ. In particular, a Legendrian unknot<br />

γ satisfies tb(γ) < 0. This inequality would be violated by the vanishing cycle of<br />

an overtwisted disc (which has tb = 0), which proves that (S 3 , ξ0) is tight.<br />

Remark 3.21. (1) Eliashberg [25] generalised the Bennequin inequality l(γ0) ≤<br />

−χ(Σ) for transverse knots (and the corresponding inequality for the Thurston-<br />

Bennequin invariant of Legendrian knots) to arbitrary tight contact 3–manifolds.<br />

Thus, whereas Bennequin established the tightness (without that name) of the<br />

standard contact structure on S 3 by proving the inequality that bears his name,<br />

that inequality is now seen, conversely, as a consequence of tightness.<br />

(2) In [9] Bennequin denotes the positively (resp. negatively) transverse push-<br />

off of the Legendrian knot γ by γ − (resp. γ + ). This has led to some sign errors in<br />

71


the literature. Notably, the ± in Proposition 2.2.1 of [25], relating the described<br />

invariants of γ and γ ± , needs to be reversed.<br />

Corollary 3.22. The standard contact structure on R 3 is tight.<br />

Proof. This is immediate from Proposition 2.13.<br />

Here are further tests for tightness:<br />

1. A closed contact 3–manifold (M, ξ) is called symplectically fillable if<br />

there exists a compact symplectic manifold (W, ω) bounded by M such that<br />

• the restriction of ω to ξ does not vanish anywhere,<br />

• the orientation of M defined by ξ (i.e. the one for which ξ is positive)<br />

coincides with the orientation of M as boundary of the symplectic manifold<br />

(W, ω) (which is oriented by ω 2 ).<br />

We then have the following result of Eliashberg [20, Thm. 3.2.1], [22] and<br />

Gromov [62, 2.4.D ′ 2 (b)], cf. [10]:<br />

Theorem 3.23 (Eliashberg-Gromov). A symplectically fillable contact structure<br />

is tight.<br />

Example 3.24. The 4–ball D 4 ⊂ R 4 with symplectic form ω = dx1 ∧dy1 +dx2 ∧<br />

dy2 is a symplectic filling of S 3 with its standard contact structure ξ0. This gives<br />

an alternative proof of Bennequin’s theorem.<br />

2. Let ( � M, ˜ ξ) → (M, ξ) be a covering map and contactomorphism. If ( � M, ˜ ξ)<br />

is tight, then so is (M, ξ), for any overtwisted disc in (M, ξ) would lift to an<br />

overtwisted disc in ( � M, ˜ ξ).<br />

Example 3.25. The contact structures ξn, n ∈ N, on the 3–torus T 3 defined by<br />

αn = cos(nθ1)dθ2 + sin(nθ1)dθ3 = 0<br />

are tight: Lift the contact structure ξn to the universal cover R 3 of T 3 ; there the<br />

contact structure is defined by the same equation αn = 0, but now θi ∈ R instead<br />

of θi ∈ R/2πZ ∼ = S 1 . Define a diffeomorphism f of R 3 by<br />

f(x, y, z) = (y/n, z cos y + x sin y, z siny − x cos y) =: (θ1, θ2, θ3).<br />

Then f ∗ αn = dz + x dy, so the lift of ξn to R 3 is contactomorphic to the tight<br />

standard contact structure on R 3 .<br />

72


Notice that it is possible for a tight contact structure to be finitely covered by<br />

an overtwisted contact structure. The first such examples were due to S. Makar-<br />

Limanov [88]. Other examples of this kind have been found by V. Colin [18] and<br />

R. Gompf [58].<br />

3. The following theorem of H. Hofer [65] relates the dynamics of the Reeb<br />

vector field to overtwistedness.<br />

Theorem 3.26 (Hofer). Let α be a contact form on a closed 3–manifold such<br />

that the contact structure kerα is overtwisted. Then the Reeb vector field of α<br />

has at least one contractible periodic orbit.<br />

Example 3.27. The Reeb vector field Rn of the contact form αn of the preceding<br />

example is<br />

Rn = cos(nθ1)∂θ2 + sin(nθ1)∂θ3 .<br />

Thus, the orbits of Rn define constant slope foliations of the 2–tori {θ1 = const.};<br />

in particular, the periodic orbits of Rn are even homologically non-trivial. It<br />

follows, again, that the ξn are tight contact structures on T 3 . (This, admittedly,<br />

amounts to attacking starlings with rice puddings fired from catapults 5 .)<br />

3.7 Classification results<br />

In this section I summarise some of the known classification results for contact<br />

structures on 3–manifolds. By Eliashberg’s Theorem 3.19 it suffices to list the<br />

tight contact structures, up to isotopy or diffeomorphism, on a given closed 3–<br />

manifold.<br />

Theorem 3.28 (Eliashberg [24]). Any tight contact structure on S 3 is isotopic<br />

to the standard contact structure ξ0.<br />

This theorem has a remarkable application in differential topology, viz., it<br />

leads to a new proof of Cerf’s theorem [16] that any diffeomorphism of S 3 ex-<br />

tends to a diffeomorphism of the 4–ball D 4 . The idea is that the above theorem<br />

implies that any diffeomorphism of S 3 is isotopic to a contactomorphism of ξ0.<br />

Eliashberg’s technique [22] of filling by holomorphic discs can then be used to<br />

show that such a contactomorphism extends to a diffeomorphism of D 4 .<br />

5 This turn of phrase originates from [93].<br />

73


As remarked earlier (Remark 2.21), Eliashberg has also classified contact<br />

structures on R 3 . Recall that homotopy classes of 2–plane distributions on S 3<br />

are classified by π3(S 2 ) ∼ = Z. By Theorem 3.19, each of these classes contains<br />

a unique (up to isotopy) overtwisted contact structure. When a point of S 3 is<br />

removed, each of these contact structures induces one on R 3 , and Eliashberg [25]<br />

shows that they remain non-diffeomorphic there. Eliashberg shows further that,<br />

apart from this integer family of overtwisted contact structures, there is a unique<br />

tight contact structure on R 3 (the standard one), and a single overtwisted one<br />

that is ‘overtwisted at infinity’ and cannot be compactified to a contact structure<br />

on S 3 .<br />

In general, the classification of contact structures up to diffeomorphism will<br />

differ from the classification up to isotopy. For instance, on the 3–torus T 3 we<br />

have the following diffeomorphism classification due to Y. Kanda [75]:<br />

Theorem 3.29 (Kanda). Every (positive) tight contact structure on T 3 is con-<br />

tactomorphic to one of the ξn, n ∈ N, described above. For n �= m, the contact<br />

structures ξn and ξm are not contactomorphic.<br />

Giroux [54] had proved earlier that ξn for n ≥ 2 is not contactomorphic to ξ1.<br />

On the other hand, all the ξn are homotopic as 2–plane fields to {dθ1 = 0}.<br />

This shows one way how Eliashberg’s classification Theorem 3.19 for overtwisted<br />

contact structures can fail for tight contact structures:<br />

• There are tight contact structures on T 3 that are homotopic as plane fields<br />

but not contactomorphic.<br />

P. Lisca and G. Matić [82] have found examples of the same kind on homology<br />

spheres by applying Seiberg-Witten theory to Stein fillings of contact manifolds,<br />

cf. also [78].<br />

Eliashberg and L. Polterovich [31] have determined the isotopy classes of<br />

diffeomorphisms of T 3 that contain a contactomorphism of ξ1: they correspond<br />

to exactly those elements of SL(3, Z) = π0(Diff(T 3 )) that stabilise the subspace<br />

0 ⊕ Z 2 corresponding to the coordinates (θ2, θ3). In combination with Kanda’s<br />

result, this allows to give an isotopy classification of tight contact structures<br />

on T 3 . One particular consequence of the result of Eliashberg and Polterovich is<br />

the following:<br />

74


• There are tight contact structures on T 3 that are contactomorphic and<br />

homotopic as plane fields, but not isotopic (i.e. not homotopic through<br />

contact structures).<br />

Again, such examples also exist on homology spheres, as S. Akbulut and<br />

R. Matveyev [2] have shown.<br />

Another aspect of Eliashberg’s classification of overtwisted contact structures<br />

that fails to hold for tight structures is of course the existence of such a structure<br />

in every homotopy class of 2–plane fields, as is already demonstrated by the<br />

classification of contact structures on S 3 . Etnyre and K. Honda [37] have recently<br />

even found an example of a manifold – the connected some of two copies of the<br />

Poincaré sphere with opposite orientations – that does not admit any tight contact<br />

structure at all.<br />

For the classification of tight contact structures on lens spaces and T 2 –bundles<br />

over S 1 see [55], [71] and [72]. A partial classification of tight contact structures<br />

on lens spaces had been obtained earlier in [34].<br />

As further reading on 3–dimensional contact geometry I can recommend the<br />

lucid Bourbaki talk by Giroux [53]. This covers the ground up to Eliashberg’s<br />

classification of overtwisted contact structures and the uniqueness of the tight<br />

contact structure on S 3 .<br />

4 A guide to the literature<br />

In this concluding section I give some recommendations for further reading, con-<br />

centrating on those aspects of contact geometry that have not (or only briefly)<br />

been touched upon in earlier sections.<br />

Two general surveys that emphasise historical matters and describe the de-<br />

velopment of contact geometry from some of its earliest origins are the one by<br />

Lutz [87] and one by the present author [45].<br />

One aspect of contact geometry that I have neglected in these notes is the<br />

Riemannian geometry of contact manifolds (leading, for instance, to Sasakian<br />

geometry). The survey by Lutz has some material on that; D. Blair [11] has<br />

recently published a monograph on the topic.<br />

There have also been various ideas for defining interesting families of contact<br />

structures. Again, the survey by Lutz has something to say on that; one such<br />

75


concept that has exhibited very intriguing ramifications – if this commercial break<br />

be permitted – was introduced in [48].<br />

4.1 Dimension 3<br />

As mentioned earlier, Chapter 8 in [1] is in parts complementary to the present<br />

notes and has some material on surfaces in contact 3–manifolds. Other surveys<br />

and introductory texts on 3–dimensional contact geometry are the introductory<br />

lectures by Etnyre [35] and the Bourbaki talk by Giroux [52]. Good places to<br />

start further reading are two papers by Eliashberg: [24] for the classification of<br />

tight contact structures and [26] for knots in contact 3–manifolds. Concerning<br />

the latter, there is also a chapter by Etnyre [36] in a companion Handbook and an<br />

article by Etnyre and Honda [38] with an extensive introduction to that subject.<br />

The surveys [20] and [27] by Eliashberg are more general in scope, but also<br />

contain material about contact 3–manifolds.<br />

3–dimensional contact topology has now become a fairly wide arena; apart<br />

from the work of Eliashberg, Giroux, Etnyre-Honda and others described earlier,<br />

I should also mention the results of Colin, who has, for instance, shown that<br />

surgery of index one (in particular: taking the connected sum) on a tight contact<br />

3–manifold leads again to a tight contact structure [17].<br />

Finally, Etnyre and L. Ng [40] have compiled a useful list of problems in<br />

3–dimensional contact topology.<br />

4.2 Higher dimensions<br />

The article [46] by the present author contains a survey of what was known at the<br />

time of writing about the existence of contact structures on higher-dimensional<br />

manifolds. One of the most important techniques for constructing contact mani-<br />

folds in higher dimensions is the so-called contact surgery along isotropic spheres<br />

developed by Eliashberg [23] and A. Weinstein [105]. The latter is a very readable<br />

paper. For a recent application of this technique see [49]. Other constructions<br />

of contact manifolds (branched covers, gluing along codimension 2 contact sub-<br />

manifolds) are described in my paper [43].<br />

Odd-dimensional tori are of course amongst the manifolds with the simplest<br />

global description, but they do not easily lend themselves to the construction of<br />

contact structures. In [86] Lutz found a contact structure on T 5 ; since then it has<br />

been one of the prize questions in contact geometry to find a contact structure on<br />

76


higher-dimensional tori. That prize, as it were, recently went to F. Bourgeois [13],<br />

who showed that indeed all odd-dimensional tori do admit a contact structure.<br />

His construction uses the result of Giroux and Mohsen [56], [57] about open book<br />

decompositions adapted to contact structures in conjunction with the original<br />

proof of Lutz. With the help of the branched cover theorem described in [43] one<br />

can conclude further that every manifold of the form M × Σ with M a contact<br />

manifold and Σ a surface of genus at least 1 admits a contact structure.<br />

Concerning the classification of contact structures in higher dimensions, the<br />

first steps have been taken by Eliashberg [28] with the development of con-<br />

tact homology, which has been taken further in [29]. This has been used by<br />

I. Ustilovsky [102] to show that on S 4n+1 there exist infinitely many non-isomor-<br />

phic contact structures that are homotopically equivalent (in the sense that they<br />

induce the same almost contact structure, i.e. reduction of the structure group<br />

of TS 4n+1 to 1 × U(2n)). Earlier results in this direction can be found in [44] in<br />

the context of various applications of contact surgery.<br />

4.3 Symplectic fillings<br />

A survey on the various types of symplectic fillings of contact manifolds is given<br />

by Etnyre [33], cf. also the survey by Bennequin [10]. Etnyre and Honda [39]<br />

have recently shown that certain Seifert fibred 3–manifolds M admit tight con-<br />

tact structures ξ that are not symplectically semi-fillable, i.e. there is no sym-<br />

plectic filling W of (M, ξ) even if W is allowed to have other contact boundary<br />

components. That paper also contains a good update on the general question of<br />

symplectic fillability.<br />

A related question is whether symplectic manifolds can have disconnected<br />

boundary of contact type (this corresponds to a stronger notion of symplectic<br />

filling defined via a Liouville vector field transverse to the boundary and pointing<br />

outwards). For (boundary) dimension 3 this is discussed by D. McDuff [91];<br />

higher-dimensional symplectic manifolds with disconnected boundary of contact<br />

type have been constructed in [42].<br />

4.4 Dynamics of the Reeb vector field<br />

In a seminal paper, Hofer [65] applied the method of pseudo-holomorphic curves,<br />

which had been introduced to symplectic geometry by Gromov [62], to solve<br />

(for large classes of contact 3–manifolds) the so-called Weinstein conjecture [104]<br />

77


concerning the existence of periodic orbits of the Reeb vector field of a given<br />

contact form. (In fact, one of the remarkable aspects of Hofer’s work is that in<br />

many instances it shows the existence of a periodic orbit of the Reeb vector field<br />

of any contact form defining a given contact structure.) A Bourbaki talk on the<br />

state of the art around the time when Weinstein formulated the conjecture that<br />

bears his name was given by N. Desolneux-Moulis [19]; another Bourbaki talk by<br />

F. Laudenbach describes Hofer’s contribution to the problem.<br />

The textbook by Hofer and E. Zehn<strong>der</strong> [70] addresses these issues, although its<br />

main emphasis, as is clear from the title, lies more in the direction of symplectic<br />

geometry and Hamiltonian dynamics. Two surveys by Hofer [66], [67], and one<br />

by Hofer and M. Kriener [68], are more directly concerned with contact geometry.<br />

Of the three, [66] may be the best place to start, since it <strong>der</strong>ives from a course of<br />

five lectures. In collaboration with K. Wysocki and Zehn<strong>der</strong>, Hofer has expanded<br />

his initial ideas into a far-reaching project on the characterisation of contact<br />

manifolds via the dynamics of the Reeb vector field, see e.g. [69].<br />

References<br />

[1] B. Aebischer, M. Borer, M. Kälin, Ch. Leuenberger and H.M. Reimann,<br />

Symplectic <strong>Geometry</strong>, Progr. Math. 124, Birkhäuser, Basel (1994).<br />

[2] S. Akbulut and R. Matveyev, A note on contact structures, Pacific J. Math.<br />

182 (1998), 201–204.<br />

[3] D.N. Akhiezer, Homogeneous complex manifolds, in: Several Complex Vari-<br />

ables IV (S.G. Gindikin, G.M. Khenkin, eds.), Encyclopaedia Math. Sci. 10,<br />

Springer, Berlin (1990), 195–244.<br />

[4] S. Altschuler, A geometric heat flow for one-forms on three-dimensional<br />

manifolds, Illinois J. Math. 39 (1995), 98–118.<br />

[5] J.W. Alexan<strong>der</strong>, A lemma on systems of knotted curves, Proc. Nat. Acad.<br />

Sci. U.S.A. 9 (1923), 93–95.<br />

[6] V.I. Arnold, Characteristic class entering in quantization conditions, Funct.<br />

Anal. Appl. 1 (1967), 1–13.<br />

[7] V.I. Arnold, Mathematical Methods of Classical Mechanics, Grad. Texts in<br />

Math. 60, Springer, Berlin (1978).<br />

78


[8] A. Banyaga, The Structure of Classical Diffeomorphism Groups, Math. Appl.<br />

400, Kluwer, Dordrecht (1997).<br />

[9] D. Bennequin, Entrelacements et équations de Pfaff, in: IIIe Rencontre de<br />

Géométrie du Schnepfenried, vol. 1, Astérisque 107–108 (1983), 87–161.<br />

[10] D. Bennequin, Topologie symplectique, convexité holomorphe et structures<br />

de contact (d’après Y. Eliashberg, D. McDuff et al.), in: Séminaire Bourbaki,<br />

vol. 1989/90, Astérisque 189–190 (1990), 285–323.<br />

[11] D.E. Blair, Riemannian <strong>Geometry</strong> of <strong>Contact</strong> and Symplectic Manifolds,<br />

Progr. Math. 203, Birkhäuser, Basel (2002).<br />

[12] W.M. Boothby, Transitivity of the automorphisms of certain geometric struc-<br />

tures, Trans. Amer. Math. Soc. 137 (1969), 93–100.<br />

[13] F. Bourgeois, Odd dimensional tori are contact manifolds, Int. Math. Res.<br />

Notices 2002, 1571–1574.<br />

[14] G.E. Bredon, Topology and <strong>Geometry</strong>, Grad. Texts in Math. 139, Springer,<br />

Berlin (1993).<br />

[15] Th. Bröcker and K. Jänich, Einführung in die Differentialtopologie, Springer,<br />

Berlin (1973).<br />

[16] J. Cerf, Sur les difféomorphismes de la sphère de dimension trois (Γ4 = 0),<br />

Lecture Notes in Math. 53, Springer, Berlin (1968).<br />

[17] V. Colin, Chirurgies d’indice un et isotopies de sphères dans les variétés de<br />

contact tendues, C. R. Acad. Sci. Paris Sér. I Math. 324 (1997), 659–663.<br />

[18] V. Colin, Recollement de variétés de contact tendues, Bull. Soc. Math. France<br />

127 (1999), 43–69.<br />

[19] N. Desolneux-Moulis, Orbites périodiques des systèmes hamiltoniens au-<br />

tonomes (d’après Clarke, Ekeland-Lasry, Moser, Rabinowitz, Weinstein), in:<br />

Séminaire Bourbaki, vol. 1979/80, Lecture Notes in Math. 842, Springer,<br />

Berlin (1981), 156–173.<br />

79


[20] Y. Eliashberg, Three lectures on symplectic topology in Cala Gonone: Basic<br />

notions, problems and some methods, in: Conference on Differential Geo-<br />

metry and Topology (Sardinia, 1988), Rend. Sem. Fac. Sci. Univ. Cagliari<br />

58 (1988), suppl., 27–49.<br />

[21] Y. Eliashberg, Classification of overtwisted contact structures on 3–mani-<br />

folds, Invent. Math. 98 (1989), 623–637.<br />

[22] Y. Eliashberg, Filling by holomorphic discs and its applications, in: Geo-<br />

metry of Low-Dimensional Manifolds (Durham, 1989) vol. 2, London Math.<br />

Soc. Lecture Note Ser. 151, Cambridge University Press (1990), 45–67.<br />

[23] Y. Eliashberg, Topological characterization of Stein manifolds of dimension<br />

> 2, Internat. J. Math. 1 (1990), 29–46.<br />

[24] Y. Eliashberg, <strong>Contact</strong> 3–manifolds twenty years since J. Martinet’s work,<br />

Ann. Inst. Fourier (Grenoble) 42 (1992), no. 1-2, 165–192.<br />

[25] Y. Eliashberg, Classification of contact structures on R 3 , Internat. Math.<br />

Res. Notices 1993, 87–91.<br />

[26] Y. Eliashberg, Legendrian and transversal knots in tight contact 3–mani-<br />

folds, in: Topological Methods in Mo<strong>der</strong>n Mathematics (Stony Brook, 1991),<br />

Publish or Perish, Houston (1993), 171–193.<br />

[27] Y. Eliashberg, Symplectic topology in the nineties, Differential Geom. Appl.<br />

9 (1998), 59–88.<br />

[28] Y. Eliashberg, Invariants in contact topology, in: Proceedings of the Interna-<br />

tional Congress of Mathematicians (Berlin, 1998) vol. II, Doc. Math. 1998,<br />

Extra Vol. II, 327–338.<br />

[29] Y. Eliashberg, A. Givental and H. Hofer, Introduction to symplectic field<br />

theory, Geom. Funct. Anal. 2000, Special Volume, Part II, 560–673.<br />

[30] Y. Eliashberg and N. Mishachev, Introduction to the h-principle, Grad. Stud.<br />

Math. 48, American Mathematical Society, Providence (2002).<br />

[31] Y. Eliashberg and L. Polterovich, New applications of Luttinger’s surgery,<br />

Comment. Math. Helv. 69 (1994), 512–522.<br />

80


[32] Y. Eliashberg and W.P. Thurston, Confoliations, Univ. Lecture Ser. 13,<br />

American Mathematical Society, Providence (1998).<br />

[33] J.B. Etnyre, Symplectic convexity in low-dimensional topology, Topology<br />

Appl. 88 (1998), 3–25.<br />

[34] J.B. Etnyre, Tight contact structures on lens spaces, Commun. Contemp.<br />

Math. 2 (2000), 559–577; erratum: ibid. 3 (2001), 649–652.<br />

[35] J.B. Etnyre, Introductory lectures on contact geometry, in: Topology and<br />

<strong>Geometry</strong> of Manifolds (Athens, GA, 2001), Proc. Sympos. Pure Math. 71,<br />

American Mathematical Society, Providence (2003), 81–107.<br />

[36] J.B. Etnyre, Legendrian and transverse knots, Handbook of Knot Theory<br />

(W. Menasco, M. Thistlethwaite, eds.), to appear.<br />

[37] J.B. Etnyre and K. Honda, On the nonexistence of tight contact structures,<br />

Ann. of Math. (2) 153 (2001), 749–766.<br />

[38] J.B. Etnyre and K. Honda, Knots and contact geometry I: Torus knots and<br />

the figure eight knot, J. Symplectic Geom. 1 (2001), 63–120.<br />

[39] J.B. Etnyre and K. Honda, Tight contact structures with no symplectic<br />

fillings, Invent. Math. 148 (2002), 609–626.<br />

[40] J.B. Etnyre and L.L. Ng, Problems in low dimensional contact topology, in:<br />

Topology and <strong>Geometry</strong> of Manifolds (Athens, GA, 2001), Proc. Sympos.<br />

Pure Math. 71, American Mathematical Society, Providence (2003), 337–<br />

357.<br />

[41] H. Freudenthal, Zum Hopfschen Umkehrhomomorphismus, Ann. of Math.<br />

(2) 38 (1937), 847–853.<br />

[42] H. Geiges, Symplectic manifolds with disconnected boundary of contact type,<br />

Internat. Math. Res. Notices 1994, 23–30.<br />

[43] H. Geiges, Constructions of contact manifolds, Math. Proc. Cambridge Phi-<br />

los. Soc. 121 (1997), 455–464.<br />

[44] H. Geiges, Applications of contact surgery, Topology 36 (1997), 1193–1220.<br />

81


[45] H. Geiges, A brief history of contact geometry and topology, Expo. Math.<br />

19 (2001), 25–53.<br />

[46] H. Geiges, <strong>Contact</strong> topology in dimension greater than three, in: European<br />

Congress of Mathematics (Barcelona, 2000) vol. 2, Progress in Math. 202,<br />

Birkhäuser, Basel (2001), 535–545.<br />

[47] H. Geiges, h-Principles and flexibility in geometry, Mem. Amer. Math. Soc.<br />

164 (2003), no. 779.<br />

[48] H. Geiges and J. Gonzalo, <strong>Contact</strong> geometry and complex surfaces, Invent.<br />

Math. 121 (1995), 147–209.<br />

[49] H. Geiges and C.B. Thomas, <strong>Contact</strong> structures, equivariant spin bordism,<br />

and periodic fundamental groups, Math. Ann. 320 (2001), 685–708.<br />

[50] V.L. Ginzburg, On closed characteristics of 2–forms, Ph.D. Thesis, Berkeley<br />

(1990).<br />

[51] V.L. Ginzburg, Calculation of contact and symplectic cobordism groups,<br />

Topology 31 (1992), 767–773.<br />

[52] E. Giroux, Convexité en topologie de contact, Comment. Math. Helv. 66<br />

(1991), 637–677.<br />

[53] E. Giroux, Topologie de contact en dimension 3 (autour des travaux de Yakov<br />

Eliashberg), in: Séminaire Bourbaki, vol. 1992/93, Astérisque 216 (1993),<br />

7–33.<br />

[54] E. Giroux, Une structure de contact, même tendue, est plus ou moins tordue,<br />

Ann. Sci. École Norm. Sup. (4) 27 (1994), 697–705.<br />

[55] E. Giroux, Structures de contact en dimension trois et bifurcations des feuil-<br />

letages de surfaces, Invent. Math. 141 (2000), 615–689.<br />

[56] E. Giroux, Géométrie de contact: de la dimension trois vers les dimensions<br />

supérieures, in: Proceedings of the International Congress of Mathematicians<br />

(Beijing, 2002) vol. II, Higher Education Press, Beijing (2002), 405–414.<br />

[57] E. Giroux and J.-P. Mohsen, Structures de contact et fibrations symplec-<br />

tiques au-dessus du cercle, in preparation.<br />

82


[58] R.E. Gompf, Handlebody construction of Stein surfaces, Ann. of Math. (2)<br />

148 (1998), 619–693.<br />

[59] J. Gonzalo, Branched covers and contact structures, Proc. Amer. Math. Soc.<br />

101 (1987), 347–352.<br />

[60] D.H. Gottlieb, Partial transfers, in: Geometric Applications of Homotopy<br />

Theory I, Lecture Notes in Math. 657, Springer, Berlin (1978), 255–266.<br />

[61] J.W. Gray, Some global properties of contact structures, Ann. of Math. (2)<br />

69 (1959), 421–450.<br />

[62] M. Gromov, Pseudoholomorphic curves in symplectic manifolds, Invent.<br />

Math. 82 (1985), 307–347.<br />

[63] H.M. Hilden, J.M. Montesinos and T. Thickstun, Closed oriented 3–<br />

manifolds as 3–fold branched coverings of S 3 of special type, Pacific J. Math.<br />

65 (1976), 65–76.<br />

[64] M.W. Hirsch, Differential Topology, Grad. Texts in Math. 33, Springer,<br />

Berlin (1976).<br />

[65] H. Hofer, Pseudoholomorphic curves in symplectizations with applications to<br />

the Weinstein conjecture in dimension 3, Invent. Math. 114 (1993), 515–563.<br />

[66] H. Hofer, Holomorphic curves and dynamics in dimension three, in: Sym-<br />

plectic <strong>Geometry</strong> and Topology (Park City, 1997), IAS/Park City Math. Ser.<br />

7, American Mathematical Society, Providence (1999), 35–101.<br />

[67] H. Hofer, Holomorphic curves and real three-dimensional dynamics, Geom.<br />

Funct. Anal. 2000, Special Volume, Part II, 674–704.<br />

[68] H. Hofer and M. Kriener, Holomorphic curves in contact dynamics, in: Dif-<br />

ferential Equations (La Pietra, 1996), Proc. Sympos. Pure Math. 65, Amer-<br />

ican Mathematical Society, Providence (1999), 77–131.<br />

[69] H. Hofer, K. Wysocki and E. Zehn<strong>der</strong>, A characterisation of the tight three-<br />

sphere, Duke Math. J. 81 (1995), 159–226; erratum: ibid. 89 (1997), 603–<br />

617.<br />

[70] H. Hofer and E. Zehn<strong>der</strong>, Symplectic Invariants and Hamiltonian Dynamics<br />

Birkhäuser, Basel (1994).<br />

83


[71] K. Honda, On the classification of tight contact structures I, Geom. Topol.<br />

4 (2000), 309–368.<br />

[72] K. Honda, On the classification of tight contact structures II, J. Differential<br />

Geom. 55 (2000), 83–143.<br />

[73] H. Hopf, Zur Algebra <strong>der</strong> Abbildungen von Mannigfaltigkeiten, J. Reine<br />

Angew. Math. 105 (1930), 71–88.<br />

[74] D. Husemoller, Fibre Bundles (3rd edition), Grad. Texts in Math. 20,<br />

Springer, Berlin (1994).<br />

[75] Y. Kanda, The classification of tight contact structures on the 3–torus,<br />

Comm. Anal. Geom. 5 (1997), 413–438.<br />

[76] R.C. Kirby, The Topology of 4–Manifolds, Lecture Notes in Math. 1374,<br />

Springer, Berlin (1989).<br />

[77] A.A. Kosinski, Differential Manifolds, Academic Press, Boston (1993).<br />

[78] P.B. Kronheimer and T.S. Mrowka, Monopoles and contact structures, In-<br />

vent. Math. 130 (1997), 209–255.<br />

[79] F. Laudenbach, Orbites périodiques et courbes pseudo-holomorphes, appli-<br />

cation à la conjecture de Weinstein en dimension 3 (d’après H. Hofer et al.),<br />

in: Séminaire Bourbaki, vol. 1993/1994, Astérisque 227 (1995), 309–333.<br />

[80] P. Libermann and C.-M. Marle, Symplectic <strong>Geometry</strong> and Analytical Me-<br />

chanics, Math. Appl. 35, Reidel, Dordrecht (1987).<br />

[81] W.B.R. Lickorish, A representation of orientable combinatorial 3–manifolds,<br />

Ann. of Math. (2) 76 (1962), 531–540.<br />

[82] P. Lisca and G. Matić, Tight contact structures and Seiberg-Witten invari-<br />

ants, Invent. Math. 129 (1997), 509–525.<br />

[83] R. Lutz, Sur l’existence de certaines formes différentielles remarquables sur<br />

la sphère S 3 , C. R. Acad. Sci. Paris Sér. A–B 270 (1970), A1597–A1599.<br />

[84] R. Lutz, Sur quelques propriétés des formes differentielles en dimension trois,<br />

Thèse, Strasbourg (1971).<br />

84


[85] R. Lutz, Structures de contact sur les fibrés principaux en cercles de dimen-<br />

sion trois, Ann. Inst. Fourier (Grenoble), 27 (1977), no. 3, 1–15.<br />

[86] R. Lutz, Sur la géométrie des structures de contact invariantes, Ann. Inst.<br />

Fourier (Grenoble) 29 (1979), no. 1, 283–306.<br />

[87] R. Lutz, Quelques remarques historiques et prospectives sur la géométrie de<br />

contact, in: Conference on Differential <strong>Geometry</strong> and Topology (Sardinia,<br />

1988), Rend. Sem. Fac. Sci. Univ. Cagliari 58 (1988), suppl., 361–393.<br />

[88] S. Makar-Limanov, Tight contact structures on solid tori, Trans. Amer.<br />

Math. Soc. 350 (1998), 1013–1044.<br />

[89] J. Martinet, Sur les singularités des formes différentielles, Ann Inst. Fourier<br />

(Grenoble) 20 (1970) no. 1, 95–178.<br />

[90] J. Martinet, Formes de contact sur les variétés de dimension 3, in: Proc.<br />

Liverpool Singularities Sympos. II, Lecture Notes in Math. 209, Springer,<br />

Berlin (1971), 142–163.<br />

[91] D. McDuff, Symplectic manifolds with contact type boundaries, Invent.<br />

Math. 103 (1991), 651–671.<br />

[92] D. McDuff and D. Salamon, Introduction to Symplectic Topology (2nd edi-<br />

tion), Oxford University Press, 1998.<br />

[93] S. Milligan, The Starlings, in: The Goon Show, Series 4, Special Programme,<br />

broadcast 31st August 1954.<br />

[94] J.W. Milnor, Topology from the Differentiable Viewpoint, The University<br />

Press of Virginia, Charlottesville (1965).<br />

[95] J.W. Milnor and J.D. Stasheff, Characteristic Classes, Ann. of Math. Studies<br />

76, Princeton University Press, Princeton (1974).<br />

[96] J. Moser, On the volume elements on a manifold, Trans. Amer. Math. Soc.<br />

120 (1965), 286–294.<br />

[97] D. Rolfsen, Knots and Links, Publish or Perish, Houston (1976).<br />

[98] N. Saveliev, Lectures on the Topology of 3–Manifolds, de Gruyter, Berlin<br />

(1999).<br />

85


[99] N. Steenrod, The Topology of Fibre Bundles, Princeton University Press,<br />

Princeton (1951).<br />

[100] I. Tamura, Topology of Foliations: An Introduction, Transl. Math. Monogr.<br />

97, American Mathematical Society, Providence, 1992.<br />

[101] W.P. Thurston and H.E. Winkelnkemper, On the existence of contact forms,<br />

Proc. Amer. Math. Soc. 52 (1975), 345–347.<br />

[102] I. Ustilovsky, Infinitely many contact structures on S 4m+1 , Internat. Math.<br />

Res. Notices 1999, 781–791.<br />

[103] A.H. Wallace, Modifications and cobounding manifolds, Canad. J. Math.<br />

12 (1960), 503–528.<br />

[104] A. Weinstein, On the hypotheses of Rabinowitz’ periodic orbit theorem, J.<br />

Differential Equations 33 (1979), 353–358.<br />

[105] A. Weinstein, <strong>Contact</strong> surgery and symplectic handlebodies, Hokkaido<br />

Math. J. 20 (1991), 241–251.<br />

86

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!